Book: Introduction to Supply Chain (Full Text)
Rough full-text extract of the book Introduction to Supply Chain (September 2025) by Joannes Vermorel.
This page is intended for full-text search. For a better looking version, check this nicer preview.
Primer
Between a bad economist and a good one lies the whole distinction: the former stops at the visible effect; the latter reckons both with the effect that is seen and with those that must be foreseen. This difference is vast [..] the bad economist pursues a small present good that will be followed by a great future harm, whereas the true economist seeks a great future good at the risk of a slight present hardship. What is Seen and What is Not Seen (1850), Frédéric Bastiat
Supply chain is an abstraction—an intellectual tool that coordinates the efforts of many to deliver physical goods cheaply and reliably. This instrument was devised to tame the incredible complexity that our industrialized civilization entails. Supply chain cannot be seen or touched. Factories, warehouses, vessels, and trucks—visible as they are—are not the supply chain. The supply chain lies in the web of expectations binding those assets together. Supply chain is an intent, not a thing.
Over the last two decades, automation has become a central concern for supply chain. If the 20^th^ century was the century of mechanizing physical labor, the 21st is decidedly about mechanizing intellectual labor. The ambient complexity of most large supply chains already vastly exceeds what can be handled even by the most diligent managers. As such, supply chain must be approached and understood in ways that maximally leverage what modern computers offer.
Approaching supply chain is an arduous task. It is at once abstract—demanding reflection on concepts like time or human nature—and concrete, requiring attention to mundane details such as a container’s maximum tonnage or national holidays. Most difficult of all, as Bastiat rightly noted, is the ability to envision what cannot be seen—only foreseen. Indeed, the connection between economics and supply chain runs deep.
We therefore need a precise definition of supply chain—a foundation that will guide the chapters ahead.
Supply chain is both a field of study and a corporate practice. The field of study is descriptive: it describes the flow of physical goods by uncovering its intrinsic principles and properties. The goal is to develop a predictive theory that assesses both the immediate and long-term consequences of every factor influencing the flow. Some levers, such as issuing a purchase order to a supplier, are within the company’s control, whereas others, like receiving a purchase order from a client, are not. Ultimately, these consequences shape the company’s long-term profitability.
The corporate practice is prescriptive. It aims to design the organization and methods best suited to guide the flow of physical goods within—and potentially beyond—the company. The goal is to pinpoint the perspectives, tools, and practices generally considered desirable for enhancing supply chains. Conversely, practices generally detrimental to profitability should be recognized as such. Naturally, the prescriptive approach is grounded in descriptive analysis: understanding decision consequences should guide the company toward better choices.
Decisions
A decision allocates a scarce resource to a specific purpose, thereby excluding alternatives that were previously available. For example, if a company spends money to purchase one additional unit of a product, that money can no longer be used for any alternative use. Over time, the company will need funds for new purchases—possibly recovered by disposing of or repurposing the previously acquired product. For now, the money remains committed as a result of the decision. Supply-chain decisions are the subset of economic decisions concerned with the flow of goods; many economic choices involve no such flow at all.
All resources, not just money, are scarce. Inventory is another fundamental resource in supply-chain settings. For example, if a unit in the warehouse is dispatched to one store, it cannot be dispatched to any other. Eventually, the company might restock the same product in the warehouse or even return the original unit from the store. However, at least for now, that inventory unit is committed as a result of the decision.
Pricing decisions also play a critical role in supply chain. Setting a product’s price creates a market condition for its exchange, effectively precluding other pricing options for that sale. Although market demand always involves uncertainty, the chosen price communicates the company’s valuation of its inventory. Over time, companies may adjust their pricing strategies for subsequent sales in response to market feedback.
Any deliberate action that alters what moves, where, when, or how much is a supply-chain decision. Raising a purchase order, scheduling a truck, allocating an inbound container to a warehouse rather than a cross-dock, or discounting a slow-mover all reshape the material flow within operational lead times; they therefore fall squarely under the remit of supply chain.
Conversely, initiatives whose influence is only remote—funding basic R&D, commissioning an advertising campaign, scouting a new geography—may eventually affect the flow but do so through several intermediate steps. Their relevance is strategic rather than operational, and thus they sit outside the scope addressed here. In practice, the boundary is crisp: among the hundreds of levers a planner manipulates every week, only a handful resist immediate classification, and for those rare cases the label—supply chain or not—has negligible impact on profitability.
What usually obstructs this recognition is not analytical difficulty but corporate tradition. Over decades, companies have labeled certain levers—pricing, merchandising, assortment choices—as the exclusive province of marketing or sales and have thereby exiled them from supply-chain conversations. Yet these very levers mold demand and, by extension, every downstream movement of goods. In the pages that follow, they will be treated where logic places them: squarely within the perimeter of supply chain.
Optionality
A decision is made among competing options. However, having multiple available options is not assured for a company. For example, if a company has identified only one supplier for a product, its choices are limited to selecting the order date and quantity from that supplier. If that supplier proves too unreliable or expensive, the company may be forced to stop ordering, likely frustrating both the business and its clients. Thus, fostering favorable options is crucial. Optionality is vital because any forecasts a company produces inherently involve irreducible uncertainty, regardless of the technology used.
Supply chain options are limited by the capacity—and, by extension, the cost—required to establish them. Management attention—and organizational attention in general—is another scarce resource that must be allocated carefully. Having multiple reliable suppliers is preferable, but it requires identifying, qualifying, and contracting them to enable timely ordering. Depending on the circumstances, this attention may be better invested in deepening the relationship with the current—and possibly best—supplier.
Viewing unused options as mere overhead to be trimmed by management is naive and misguided. There may be cases where a company overinvests in optionality, yielding diminishing returns when additional options fail to deliver proportional benefits. Nonetheless, such overinvestments are infrequent, even if maintaining options may seem costly to outsiders.
Numerous options are self-evident, yet for practical purposes many are ruled out. This happens whenever options cannot be easily assessed. For instance, a retail network with many stores may recognize occasional opportunities to rebalance inventory by selecting an optimal route for a truck to load and unload goods sequentially. Yet, if such rebalancing must be planned by hand—or with generic tools ill-suited to the task—the combinatorial explosion swiftly overwhelms people and spreadsheets. In the absence of dedicated optimization software able to sift through millions of routing permutations in seconds, this class of options is quietly dismissed as “too complex” and never reaches the decision table.
More often, options are assessed only intermittently, largely due to the company’s limited attention or limited tooling. For example, a company might place orders with a supplier only once a week—not because weekly cycles have inherent limitations or benefits, but simply because composing daily orders would be too time-consuming. A daily assessment does not mean an order is placed every day; it means the option to order is reviewed daily, even if no order is made on most days.
Physical goods
Supply chain is about the flow of physical goods. Remove the goods, and what remains is little more than finance and trading. The craft unfolds in three recurring moves: first, money is converted into goods—stock units purchased from suppliers; next, those goods are processed—inspected, stored, transported, or transformed; finally, the goods are turned back into money, whether through direct sale or by being consumed in producing further products and services.
The process is meant to be repeated indefinitely, though never in the exact same way, as every aspect of the flow changes over time. Only the existence of the flow remains constant, enduring as long as the company exists.
Physical goods impose constraints that cannot be eluded. Every action involving them—purchasing, moving, transforming—burns cash, yet every non-action—keeping, waiting, storing—levies an implicit toll just the same. Money is spent whether the goods flow or stand still; the only leverage lies in choosing the mix of actions and inactions that minimizes total cost while preserving optionality. These realities remain invisible when goods are reduced to electronic markers in modern trading systems. The high-frequency trader dealing in copper need not care where the metal rests, how it is protected from oxidation or theft, or whether it fits the gross-weight limit of a twenty-foot container. The supply-chain practitioner, conversely, must grapple with each of these constraints, and the market rewards companies that master both the explicit costs of action and the implicit costs of inaction.
Handling physical goods is constrained by matter, energy, space, and time. Crates cannot tunnel through walls, pallets cannot out-lift their forklifts, and a vessel that sailed from Shanghai yesterday will not berth in Rotterdam tomorrow. By contrast, the informational layer that steers those movements is, for practical purposes, elastic: a commodity laptop can explore millions of permutations per second, and a modest cloud cluster can tame billions.
Elastic does not mean trivial. Turning raw data into actionable insight is fiendishly difficult—noise, ambiguity, and sheer volume conspire to blur the picture. Yet every minute a planner spends waiting for a report to refresh, every replenishment batch delayed by “system limitations”, flag a design failure. The bottleneck should reside in steel and diesel, not in an SQL query.
Timing makes the point vivid. An algorithm that computes procurement recommendations for a thousand SKUs in fifty milliseconds is instantaneous next to the fortnight needed to manufacture the goods and the six weeks needed to float them across an ocean. Relative to atoms, bits move at light speed; any decision latency that looms large beside physical lead time is unnecessary waste.
Scale tells the same story. Enumerating ten million cartons costs pennies in storage and milliseconds of CPU. Housing those cartons demands hectares of racking, fleets of conveyors, and armies of operators. When a project is declared “blocked by data volumes”, the remedy is never more forklifts; it is always better data modeling and leaner software.
In short, the limiting factor of a well-run supply chain ought to be reality itself—cubic meters, kilowatts, calendar days. Whenever the choke point sits in the digital cockpit—dashboards that stall, optimizers that time-out, planners who “cannot see” the options—management has located a low-hanging opportunity. Eliminating decision friction is not cosmetic; it is the primary lever through which modern computing squeezes more returns from the same scarce resources.
Variability
Every supply-chain decision is a wager on the state of the firm and its market at the moment the goods finally move. Because shifting or transforming atoms always takes time—sometimes hours, often days—the decision must reach into a future that is fundamentally uncertain. That future may mirror today’s assumptions, or it may diverge sharply, turning a careful plan into a loss or, with equal ease, into an unexpected windfall.
For example, when a company places an order with an overseas supplier, it must navigate multiple uncertainties. Production and transportation delays are inherently unpredictable, as neither the manufacturer nor the carrier is entirely reliable. Furthermore, the quantity of goods received might be diminished due to random transit damage. Finally, customers’ future willingness to pay for these goods remains uncertain—further complicated by the company’s lack of influence over competitors’ pricing strategies during and after delivery.
A brief comparison between the isolated manufacturing process and supply chain as a whole clarifies the nature of variability. On the factory floor, any deviation from the nominal thickness of a steel sheet, the purity of a reagent, or the timing of a pick-and-place arm is a defect. Variability is an enemy that can always be beaten—at a price—and the price keeps falling as sensors, actuators, and algorithms push control ever closer to the atomic scale; the semiconductor industry is proof enough. Given sufficient capital, the residual spread around the target mean can be driven arbitrarily low. In supply chain, the situation is inverted: demand, lead times, supplier reliability, and competitive moves are produced by countless independent agents whose behavior cannot be engineered away. No amount of money can purchase a deterministic future. The only rational stance is to expect variability, cultivate options, and design decisions that turn uncertainty into profit rather than loss.
Indeed, supply chains are exposed to a broader world on a scale far beyond that of a closed manufacturing environment. No company can shield itself entirely from external forces—be they from clients, suppliers, regulators, the military, or competitors. In addition to man-made factors, natural events—such as snowstorms, floods, forest fires, tsunamis, earthquakes, and pandemics—still disrupt supply chains, even though the human toll such catastrophes exact has, thankfully, declined markedly over the last century.
Unlike in a manufacturing process, where variability is invariably a net negative, supply chain variability can often be transformed into a competitive advantage. Consider a bottled-water producer that intentionally builds filling capacity far above its average demand. Most of the time the extra lines sit idle; yet when a heatwave strikes—an episode that recurs only a handful of summers per decade—the company can surge output overnight while competitors ration stock. The transient spike in volume, coupled with the premium consumers willingly pay under sweltering temperatures, more than repays the dormant steel. Surplus capacity here is an option bought in advance precisely to monetize extreme weather—a textbook illustration of variability turned into profit.
In short, whenever variability brings risks, it simultaneously opens up opportunities. In fact, risks and opportunities reveal more about an observer’s vested interests than about their intrinsic nature. Tour operators organizing trips to Mount Etna have a vested interest in the volcano’s continued activity. For them, unlike the people of ancient Pompeii, an active volcano is not a disaster but a crucial source of income.
Mastery
Mastering a supply chain means nurturing favorable options while committing just enough resources to remain profitable today and resilient tomorrow. Such mastery is never the feat of a lone planner; it crystallizes through the coordinated work of many disciplines acting in concert.
Even in well-aligned organizations, five stubborn traits keep supply chains hard to tame: intangibility, mediation, complexity, connectivity, and wickedness. The pages that follow examine each trait in turn and show how it can be transformed from a hindrance into a lever of competitive advantage.
All visible assets—factories, trucks, pallets, even people—matter to the supply chain only through the intent that animates them. Supply chain is thus fundamentally intangible: a set of expectations about future exchanges of goods, money, and service.
Consider a single bottle of milk resting on a supermarket shelf. Tangible though it is, its presence is justified only by a pair of overlapping wagers. The retailer bets that a customer will buy the bottle before it spoils; the customer bets that the retailer will keep the shelf stocked. These bets remain silent when they succeed and become painfully audible only when they fail—when the milk expires or the shelf is bare.
This intangibility is the first obstacle to mastery. Expectations never appear on a balance sheet; they must be inferred from their symptoms—stockouts, overstocks, lead-time drift, margin erosion. Like economists, supply-chain practitioners advance by surfacing the latent assumptions buried in every MOQ, every replenishment cycle, every promotional calendar. Mastery begins only once those assumptions are made explicit, quantified, and challenged.
Modern supply chains cannot be observed directly; every glimpse is mediated by software records. What planners read as on-hand inventory, open purchase orders, or transit quantities is only a digital representation, assembled by an applicative landscape pieced together over decades. Crucially, this landscape did not arise as a scientific experiment designed to capture the most relevant data—contrary to what many academic treatments imply—but as a patchwork of tools built to keep operations running day after day. Consequently, the landscape dictates both the visibility and the vocabulary of the flow; the atoms themselves remain one step removed.
Practitioners often worry about numerical inaccuracies—an errant cycle
count, a late barcode scan—yet such errors are minor and usually
contained. The real hazard is semantic drift. A field that meant
confirmed_order
yesterday may, after a vendor patch or an improvised
user workaround, capture something quite different tomorrow while its
label stays unchanged. Meanings mutate faster than counts, and
distortions multiply as data hops between systems. Each record is
therefore a hypothesis about reality, not reality itself, and must be
revalidated continuously.
The complexity of supply chains far surpasses the capacity of any single mind, regardless of the effort invested. Even modest firms now juggle hundreds of thousands of SKUs (stock-keeping units). Part of this proliferation is accidental—legacy regulations, haphazard processes, badly patched software. Yet an ever-larger share is intentional: intricacy deliberately introduced because it augments the value proposition.
E-commerce makes the point vivid. Delivering a toothbrush to a studio apartment entails last-mile routing, parcel tracking, and reverse logistics—steps absent from a pallet drop at a hypermarket. The extra layers are undeniably complex, yet they spare customers the trip and therefore command a premium. In aviation, engine makers no longer sell hardware; they sell “power by the hour”, aligning perfectly with what airlines value while shifting maintenance variability onto the manufacturer’s MRO network. Likewise, a fashion brand that broadens its assortment from 1,000 to 10,000 SKUs complicates buying, stocking, and markdowns, but the wider choice enlarges the addressable audience and cements loyalty.
Such essential complexity earns its keep because it creates customer value that can be monetized—provided it is profitably tamed. What remains—compliance quirks, phantom SKUs, policy relics—is accidental complexity and deserves to be excised. Distinguishing the two, suppressing the accidental, and harnessing the essential is one of the grand challenges of supply chain in the 21^st^ century.
Supply chains operate as systems in the sense articulated by Russell Ackoff: the performance of the whole emerges from the interactions of its parts, not from the isolated excellence of each. They are therefore inherently interconnected. Every decision—however small—ripples through the entire network. Consider a single unit of inventory resting in a regional warehouse. The moment a planner allocates it to Store A, all other stores are instantly deprived of that option. A local relief—preventing a stockout at Store A—may simultaneously seed future shortages at Stores B, C and D. This zero-sum tension is structural, for every product, supplier, truck, and minute of managerial attention draws upon the same finite pools of resources.
These couplings are magnified by economic forces. Customers expect complementary items to be available together; carriers quote full-truck rates, not pallet rates; a conveyor belt amortizes its capital only when volume is consolidated. Attempting to “optimize” one node in isolation merely displaces cost or risk elsewhere in the chain. Such mechanical, divide-and-conquer reasoning—excellent for machines—fails for supply chains because it ignores the feedback loops that bind decisions across time and space.
Mastery therefore demands systems thinking: evaluating every lever through its influence on the entire flow, both now and in the future. Only a company that sees warehouse allocation, truck scheduling, replenishment rules, and pricing as facets of the same decision space can transform interdependency from a handicap into an advantage.
The problems that supply chains face are wicked—not in the moral sense, but in the technical sense coined by Rittel and Webber to describe challenges that mutate as soon as one intervenes. A wicked problem has no crisp formulation, no final solution, and every “fix” reshapes the very system it tries to mend.
A concrete illustration makes this plain. Suppose a company announces a bonus tied to “94 % forecast accuracy.” Once planners’ livelihoods depend on that figure, the fastest route is to shorten horizons, lump SKUs together, or pad safety stocks until the arithmetic behaves. Working capital soars, service slides, yet the dashboard glows green. The measure became the target and thus ceased to be a good measure—Goodhart’s Law laid bare.
The paradox runs deep. Supply-chain management needs explicit, numerical targets so optimizers—often software immune to personal agendas—can pursue them relentlessly. Yet the moment a target is announced, it broadcasts fresh incentives for employees, suppliers, and even competitors to game the rules. No prescription remains optimal for long, and any rule that is obvious, static, or predictable will be undermined as agents adapt. Wickedness therefore overturns naive academic hopes of tidy “problem + resolution” pairs: mastery lies in continually revisiting both the objectives and the algorithms that chase them.
Parting thoughts
The deceptively compact definition of supply chain hides a radical prospect. Once every lever that moves goods is captured as data, decisions no longer wait for a planner to click “OK”; they can be enacted, audited, and refined by algorithms that run day and night at negligible marginal cost. Automation—understood as unattended, software-driven decision making—is poised to do for white-collar schedulers what the assembly line did for manual labor.
Far from being mere labor arbitrage, such automation unlocks productivity that spreadsheet-ridden organizations can scarcely imagine. Many companies now employ more analysts grooming spreadsheets than operators handling cartons, yet a single decision engine can outpace both, slashing latency and error while freeing scarce talent to tackle novel, higher-order challenges.
More importantly, software can roam decision spaces that humans never will. Continuous repricing during viral demand spikes, mid-voyage rerouting of ocean containers, or continent-wide inventory balancing every hour becomes tangible once computation—not human patience—is the bottleneck. Just as e-commerce reinvented retail rather than polishing telesales, unattended optimization will reshape supply chains instead of merely accelerating yesterday’s procedures.
The quantitative supply-chain literature amassed since the UNIVAC I (1951) overflows with intimidating algebra and petabyte-hungry algorithms, yet it still fails to harness modern computing to deliver tangible supply chain results. Its sophistication is misdirected: it optimizes what is easy to tally rather than what is economically decisive, clings to brittle point forecasts, and collapses when confronted with the combinatorial sprawl of real-world options. By missing the target, it turns today’s near-free gigaflops into background noise while planners continue to wrestle with spreadsheets. The qualitative prescriptions on leadership have fared no better: libraries are full of supply-chain “handbooks” that few leaders read and from which fewer still would benefit.
A striking symptom of this vacuum is the prevalence of self-taught practitioners in large companies. Society rightfully tolerates no self-taught surgeons—the gap between amateurs and trained experts is too vast to risk human life. Yet when billions in inventory are at stake, enterprises routinely rely on planners who have pieced together their craft from tribal lore and blog posts. The comparison is stark, and it underscores how underdeveloped the discipline remains.
Mainstream methods have not merely underdelivered; in many cases they have actively misled, promoting illusory precision and spreadsheet heroics in place of scalable science. Their shortcomings—and the path to move beyond them—are examined head-on in the final chapter of this book.
The chapters ahead assemble the scientific building blocks needed to encode supply-chain mastery into software and to unlock the unattended optimization that tomorrow’s leaders will take for granted. Replacing folklore with a precise, computer-ready science is the prerequisite for turning variability and optionality from recurring headaches into repeatable gains.
History
Definitions of logistics, supply chain, and supply chain management (SCM) diverge: some restrict “supply chain” to the physical flow of goods and reserve “SCM” for coordination; others treat the terms as synonyms. Vendors, consultants, and practitioners twist the terms further to flatter org charts and marketing pitches; a brief historical detour helps explain the resulting fragmentation.
Supply-chain ideas sit between two poles: Euclidean crispness and AI’s shape-shifting jargon. Labels change every decade; the underlying task—mastering optionality in the flow of physical goods—does not. Claims of expertise in logistics, operations research, supply chain, or enterprise optimization often reveal more about the claimant’s vantage point than about the field.
This chapter serves a dual purpose. First, it retraces how the loose, ever-shifting label supply chain emerged from earlier notions such as logistics and operations research, showing that today’s meaning is still provisional. Second, it illustrates the adversarial mindset that will thread through the rest of the book. Below, each buzzword is paired with the incentives that birthed it. Vendors trumpet novelty to sell licenses; consultants repackage yesterday’s ideas to sell slides; executives, anxious not to fall behind, adopt the language before testing the substance. Sustaining this skeptical posture protects practitioners from costly mirages while still letting authentic innovations surface.
Early 1800s to pre-war 1900s
Long before “logistics” or “supply chain” were coined, the activity itself was the lifeblood of every city-state. Rome relied on African grain fleets, Amsterdam on Baltic timber convoys, and Edo on rice barges timed to favorable tides. Each case posed the same enduring puzzle: bridging, day after day, the spatial gap between production and consumption amid uncertainty. The challenge endures; the jargon and analytical tools change. The first systematic mathematical attempts surfaced only in the late 18^th^ century—a story we now turn to.
If the modern supply chain era means pursuing operational improvements analytically, it begins in that period. In 1781, French mathematician Gaspard Monge published his Memoir on the Theory of Embankments and Rubble (Mémoire sur la théorie des déblais et des remblais), proposing to minimize the labor required to move soil for leveling terrain with both elevations and depressions.
The term logistics, once used interchangeably with supply chain, took shape in a military context during the first half of the 19^th^ century. The earliest attested use—logistique—appears in the Summary of the Art of War (Précis de l’Art de la Guerre) (1830) by the Swiss officer Antoine-Henri Jomini1. Logistics meant moving resources—people, equipment, materials—to the right place at the right time. It refined the longstanding concept of military stewardship dating to antiquity.
In 1840, the British Post Office adopted the “Uniform Penny Post”, an early civilian demonstration of what we would now call operations research. Mathematician-engineer Charles Babbage compiled cost tables showing distance contributed only marginally to postal expenses; the real drivers were collection, sorting, and final delivery. Armed with these quantitative insights, Parliament accepted a nationwide flat rate of one penny per letter, a decision that radically expanded the volume of mail and, by extension, the reach of commerce. Babbage’s study prefigured modern supply-chain analytics by replacing folklore with measurement and optimization. His later pursuit of programmable calculating machines—culminating in the Difference Engine—aimed to automate exactly the statistical tabulations his postal work required.
In 1888, economist Francis Ysidro Edgeworth, of Irish and Spanish descent, articulated an insight later known as “the newsvendor problem”. His idea, first introduced in his Mathematical Theory of Banking, described the challenge of determining the optimal inventory level when excess stock must be disposed of at the end of a period, as is done daily with unsold newspapers. The contribution blended statistical risk analysis with operational problem-solving—a perspective that would not become mainstream for roughly a century.
The rise of a stock-owning middle class in the United States at the end of the 19^th^ century, in turn, gave birth to the early economic forecasters. Among these pioneers, Roger Babson and Irving Fisher rose to prominence at the start of the 20^th^ century. They promoted quantitative methods for allocating resources to individuals and corporations—a spirit echoed by many modern enterprise software offerings. They also pioneered statistical time-series forecasting techniques.
The evolution of economics in the first half of the 20^th^ century spurred numerous methods for solving small operational puzzles. Several of the attempted solutions led to theories and formulas that would become enduring supply chain shibboleths. For example, the economic order quantity (EOQ) formula—a supply-chain staple—was introduced in 1913 by Ford Harris and popularized in the 1930s as Wilson’s formula.
WWII and post-war developments
At the outbreak of the Second World War, the field returned to its military roots. Britain assembled nearly 1,000 men for what became operations research, defined by the British Army as “a scientific method of providing executive departments with a quantitative basis for decisions regarding the operations under their control”. The effort delivered resounding results: fewer artillery rounds to down enemy aircraft and higher survivability for British ships and planes.
Buoyed by its wartime prestige, operations research rapidly expanded in academic circles. The advent of general-purpose programmable computing hardware enabled the community to explore methods that would have been utterly impractical to compute by hand. Hardware limits kept most 1950s–60s research on static or stationary problems—those assumed unchanged over time. For example, determining the location of warehouses to service a retail store network, if we assume the demand to be constant, is a stationary problem. In contrast, optimizing warehouse allocation processes and dispatch routes is dynamic, as these problems are sensitive to fluctuations in demand, lead times, traffic, driver availability, truck maintenance, and more. Dynamic problems outstripped the hardware of the day and were largely set aside.
Although operations research initially captivated corporate managers, its appeal began to fade in the 1970s. By the 1980s, wartime prestige had faded and corporate interests had shifted2. The 1970s brought novel advances, including computing hardware capable of leveraging barcodes3. Coupled with Codd’s 1970 relational database, barcodes would reshape supply chains and their digital landscape in subsequent decades. This convergence yielded the inventory management system—an early precursor to modern business software.
In software circles, “management” then meant the routine tasks of lower-level managers. The so-called “management system” was designed to replace the manager entirely. Although their work was largely clerical, these managers belonged to an educated workforce adept at keeping meticulous, accurate records. Moreover, these white-collar workers held positions of authority and prestige that surpassed those of blue-collar workers responsible for physically moving goods.
With the emergence of management systems, software vendors soon recognized that supply chain management encompasses two distinct facets: bookkeeping and decision-making. Bookkeeping involves electronically tracking stock levels and handling routine clerical tasks. Decision-making involves determining the optimal inventory orders and allocations for the company. Vendors initially aimed to automate the managerial role; bookkeeping yielded readily, but decision-making proved vastly more complex—well beyond 1970s hardware.
Nevertheless, “bookkeeping” lacked the allure of “management”. Software vendors opted for the more marketable term—even at the cost of clarity—because it bolstered their product image. Subsequently, the software industry widely adopted the term “management” in naming systems: WMS (warehouse management system), CRM (customer relationship management), PLM (product lifecycle management), etc.
The ubiquity of “management” eventually diluted the prestige of the title “manager”. In the 2020s, the title “manager” frequently implies managing only oneself—as seen in roles like project manager, office manager, or account manager. Conversely, employees with genuine hierarchical authority typically benefit from job titles that are more specific than “manager”. Instead, these employees might be assigned titles like “lead” or “leader”.
Thus, “supply chain management” denotes the activities of managers in the traditional corporate sense—individuals with formal authority. In contrast, “supply chain” is broader, naming both the practice and the academic field.
1980s to early 2000s
By the 1980s, software vendors expanded their scope beyond inventory to encompass orders, payments, salaries, and more. Distributed software was not yet economically viable for vendors. Thus, if a company was to acquire one system to bookkeep inventory and another to bookkeep salaries, the later integration of the two systems was usually deemed impractical. Vendors therefore consolidated functions into a single mainframe monolith.
From post-war fame, logistics had slipped behind finance and marketing in prestige. A blunt aphorism circulating in large consumer-goods companies at the time summed up the prevailing hierarchy: “Smart people go to marketing, reliable people to production, and those lacking both qualities to logistics.” Although clearly colored by amicable rivalry between corporate vocations, the saying captured a widespread perception of diminished status within the logistics function.
One explanation is that, until the 1980s, logistics was largely empirical. The initial promise of operations research to transform logistics into a modern analytical science had faded. In 1979, Russell Ackoff, an American pioneer of the corporate era of operations research, summarized the situation in his paper The Future of Operational Research is Past:
Operations Research is dead even though it has yet to be buried. I also think there is little chance for its resurrection because there is so little understanding of the reasons for its demise […] The life of OR [Operations Research] has been a short one. It was born here late in the 1930s. By the mid-60s it had gained widespread acceptance in academic, scientific, and managerial circles. In my opinion this gain was accompanied by a loss of its pioneering spirit, its sense of mission and its innovativeness. Survival, stability and respectability took precedence over development, and its decline began.
In the 1980s, decision-making stagnated while bookkeeping flourished; by decade’s end, many companies had digitized their transactional operations. Electronic records enabled new ways to monitor operations; simple totals and averages, recalculated continuously, began to reshape corporate practice.
Having made modest early-20^th^-century debuts, consultancy firms became corporate giants by the early 21^st^ century. Among many causes, the spread of electronic records in the 1980s significantly boosted management consulting.
Despite the mundane observation that the sole occupation of consultants appears to be wrestling with spreadsheets and slides4, the corporate rationale is that consultants are in the business of selling impact. Those electronic records created numerous opportunities to enhance a variety of processes without resorting to any grand level of analytical sophistication.
Consulting’s perennial challenge is differentiation. As consulting comes with low entry barriers, it is difficult to maintain an offering that remains distinct from what the competition is offering. Thus, it is hardly surprising that influencing the evolution of corporate language is highly appealing to consultants. A now-classic move is to coin—or at least adopt—new terms every few years to refresh the offering and brand. The authority gained from being ostensibly linked with a novel, trendy buzzword creates a highly desirable competitive edge, at least until competitors inevitably replicate it.
Supply Chain Management (SCM) was coined by the British logistician-turned-consultant Keith Oliver in 1982. Sales and Operations Planning (S&OP) was coined in 1985 by American software engineer-turned-consultant Richard Ling. Both perspectives revolve around the idea of seeking end-to-end alignment of the various corporate functions. Both practices would have remained abstract without the increasing availability of digital corporate records.
Backed by software vendors, consultant-advocated practices spread and demanded new skill sets. Logistics directors became responsible for producing time-series forecasts and tuning service-level parameters, as exemplified by the safety stock formula. These new, more analytical responsibilities marked a significant departure from traditional logistics directors, who were primarily field leaders capable of managing large teams of warehouse workers and truck drivers (no small feat).
By the 1990s, large vendors integrated a bewildering array of bookkeeping processes into software monoliths. For clarity, these products should have been named ERM (‘Enterprise Resource Management’); however, this terminology never caught on. By then, computing hardware had advanced so much that the decision-making aspect was no longer considered intractable. Enterprise clients soon realized that, despite the claims, these tools were limited to bookkeeping.
Nevertheless, many vendors began—or resumed—investing in the decision-making aspects of supply chain. These vendors wanted to avoid confusion with bookkeeping competitors, so they differentiated their offerings. For this reason, vendors, assisted by market analysts, coined the term ERP (Enterprise Resource Planning). Within enterprise software, “planning” became the rallying cry signaling decision-making capability. Indeed, with a master plan that drills down to the most granular level—thanks to improved computer hardware—every decision becomes, in theory, merely a matter of clerical execution5. In practice, planning modules combined point time-series forecasting with deterministic inventory optimization. The intent was to deliver a technological solution that would replace the lower-level manager entirely in one fell swoop, rather than gradually.
Soon, competing vendors labeled their products ERPs regardless of how little planning they actually included. Granted, enterprise products are inherently complex, aiming to cover a wide array of scenarios; opacity often extends to clients and even to vendors themselves. As a result, vendors enjoyed considerable freedom in choosing which capabilities to showcase in their marketing materials. One might even argue that a simple moving average sales forecast qualifies as ‘planning’, thereby justifying the ERP label. Many vendors took advantage of this semantic legerdemain.
From the 1990s onward, new vendors focused exclusively on decision-making, positioning themselves as an analytical layer above the transactional layer offered by ERPs (which arguably should have been called ERMs). In the supply chain context, these vendors broadly fell into two camps: the algorithmic and the forensic. Both camps leveraged the same electronic records found in preexisting business systems.
The algorithmic camp focused on time-series forecasting techniques as well as the inventory optimization techniques underlying the planning paradigm. Because ERPs already claimed planning, marketing extra “planning” was difficult; emerging vendors, aided by analysts, coined “Advanced Planning System” (APS). In other words, “yes, your ERP has planning already, but our complementary solution has advanced planning.”
The forensic camp concentrated on producing business reports designed to help managers make better decisions. ERPs already featured “business reporting”, so software vendors—again with analyst help—popularized “business intelligence” (BI) for this class of tools6. Technically, BI’s multidimensional cubes marked a notable departure from Codd’s relational model two decades earlier.
Present day considerations
In 2025, the descendants of both camps (algorithmic and forensic) remain active. Within the algorithmic camp, integrated business planning is promoted with an emphasis on machine learning—pitched as a superior descendant of earlier statistical techniques. In the forensic camp, some promote concepts such as “supply chain digital twins” or “supply chain control towers,” which are aimed at providing insights to company leadership.
All this shows that creatively rebranding existing products with the latest buzzwords—a practice traceable to the 1970s—remains prevalent. The case of “planning” capabilities was merely a minor skirmish in the history of enterprise software. Computing hardware and software have advanced at an impressive pace over the past 70 years. Due to continuous advances in hardware, technological obsolescence arguably poses the greatest threat to enterprise software vendors today. Newer hardware makes once-impossible feats feasible—but typically demands complete software re-engineering to exploit its capabilities. Given this reality, it is far easier—and more cost-effective—to rebrand one’s offering than to completely re-engineer it to fully leverage the latest technological developments.
Software vendors encounter a real-world example of the classic prisoner’s dilemma. Collectively, vendors would benefit from more honest and transparent communication about their products. Making promises that cannot be delivered inevitably leads to costly problems for both parties. Yet, if one vendor exaggerates its claims more than competitors, it gains a strategic market edge by making its solution more appealing to clients. Overpromising and underdelivering are thus unsurprising consequences of enterprise software’s dynamics.
The lexical inflation soon reached the org chart itself. Supply chain directors emerged in the early 2000s—sometimes as a simple rebranding of the incumbent logistics director, sometimes as a freshly minted role meant to champion cross-functional analytics. By the 2020s, logistics directors seldom engaged in sophisticated predictive analytics, whereas supply chain directors rarely handled day-to-day warehouse operations.
Although advanced analytics—machine learning and mathematical optimization—have been available for about three decades, adoption in supply-chain divisions remains the exception; spreadsheets still dominate decision-making.
Today, a logistician handles warehousing and transportation. When logistics is outsourced, providers are known as “third-party logistics providers” (3PLs), which receive an electronic order for each stock movement. Ironically, although logistics was originally founded on the principle of moving resources at the right time and place, 3PLs profit from inefficiencies—charging for inventory movement and for storage duration. Thus, the greater the number of movements and the longer the storage durations, the higher the revenue earned by the logistician.
Mainstream perception
Over the last decade, the term “supply chain” has become increasingly common in mainstream media. In news coverage, a “supply chain” refers to a network of specialized companies that cooperate—sometimes very closely—to deliver sophisticated products, such as semiconductors or automobiles. News about a “supply chain” typically emerges from negative events—such as a catastrophic port fire—that not only affect one entity but also ripple through a network of interdependent companies, amplifying the original impact.
Journalists frequently question what should be done to prevent such problems from happening again in the future. While legitimate, this perspective overlooks the absence of any central authority governing those “supply chains”. Indeed, in this context, a supply chain is composed of numerous uncoordinated companies. Furthermore, introducing an authority—such as through government intervention—would likely worsen the situation, as such an authority would lack the necessary information to allocate resources effectively. As Friedrich Hayek noted in The Use of Knowledge in Society (1945), relevant knowledge is widely dispersed among countless individuals in disparate companies. No single authority can ever fully harness this decentralized information, and any attempt to impose a “captain” inevitably leads to resource misallocation.
Such newsworthy failures are emergent properties of the economy’s division of labor. Any division of labor carries inherent fragility; if a highly specialized segment falters, its repercussions spread throughout the economy. The only alternative to this fragility is to renounce the division of labor, an unreasonable proposition. Indeed, it would dramatically increase the cost of physical goods consumed every day, from rolls of toilet paper to smartphones. The tremendous benefits of the division of labor have been well-understood for centuries, with the first formalization dating back to the 19^th^ century with the work of David Ricardo and his law of comparative advantage.
Yet the division of labor characterizes markets as well: people trade precisely because specialization benefits both sides. The “supply chains” observed by the media are merely groups of tightly dependent companies that emerge due to the competitive advantages conferred by their respective specialization.
Supply chains are also conflated with outsourcing, perceived as “stealing” jobs from wealthier Westonia to benefit poorer Ruritania. Yet, in modern settings, gaining access to cheap labor overseas is rarely more than a secondary concern. If supply chains were solely about accessing the cheapest labor, then Africa would be a prime supply chain partner for advanced nations. Yet, the opposite is true: advanced nations primarily trade between themselves. The least-developed countries—often with the cheapest labor—are largely absent from global trade.
In practice, infrastructure, technology, and capital requirements usually overshadow pure labor considerations. For instance, the containerization boom of the 1950s–60s drastically reduced maritime shipping costs, but also demanded massive port upgrades that only countries with robust capital investments and stable regulatory frameworks could afford. More recently, high-value electronics require advanced robotics, rigorous quality standards, reliable power, and specialized engineering talent—complexities that dwarf the advantage of marginally lower wages. As a result, countries with strong logistics networks, dependable legal systems, and a critical mass of technical know‑how benefit from large trade volumes with similarly developed nations, whereas regions lacking these assets often remain largely disconnected from modern supply chain opportunities.
Looking ahead
Supply-chain terminology remains in flux, and vendors are more active than ever. Some consultancies advocate the “agile”, “dynamic”, and “holistic” supply chain—and for “supply-chain design”. Some forgo the “supply chain” suffix altogether, as in flowcasting or Demand Driven Material Requirements Planning (DDMRP). All these paradigms address similar problems, although their methods and techniques differ significantly.
Given that over a million scientific papers and more than ten thousand books have been published in the last 70 years, the field of ‘supply chain’ is generally regarded as mature. Yet companies still rely on spreadsheet heuristics that bear little resemblance to mainstream supply-chain theories.
One could blame practitioner ignorance of the vast literature; the charge is hardly tenable. Most large companies with an international footprint operate under fierce competitive pressure. Given how little mainstream theory has evolved over the last half-century, claiming persistent corporate ignorance strains credulity.
A simpler proposition is that mainstream theory is inadequate: practice deviates because the theory is flawed and the anticipated profits fail to materialize. In other words, despite its vast literature, supply chain remains in a pre-scientific stage where knowledge fails to yield consistent and predictable results.
Over seventy years, bookkeeping layers leaped forward—barcode scans, real-time databases, cloud ledgers—yet the rules deciding what to buy, make, move, or stock still resemble mid-century textbooks. Progress has been spectacular on the transaction side; on the decision side it has stalled. The root is not technical but epistemic: the field lacks a robust way to separate durable insight from fashionable doctrine.
Explaining—and ultimately closing—this gap requires a pause in the historical narrative. The next chapter, therefore, shifts from chronology to epistemology. Before prescribing remedies, we will examine how claims are generated, who benefits from which narratives, and what tests ideas must pass before governing a single pallet. Only once these standards are explicit can progress in decision-making—interrupted for five decades—resume in earnest.
Epistemology
Improving a supply chain begins with improving the knowledge that steers it. This chapter therefore shifts from the mechanics of moving goods to the epistemic question that precedes every recommendation: How do we decide that an idea about supply chain merits belief, funding, and deployment? Answering this question is not idle philosophy; it is the only safeguard against the fashionable doctrines, glossy slides, and “proven” algorithms that routinely consume budgets while leaving performance unchanged.
Supply-chain knowledge is unusually error-prone. Uncertainty blurs the feedback loop between decision and outcome; incentives reward confident storytelling over careful testing; and the sheer physical latency of global flows makes controlled experiments rare. As a result, spreadsheets still dominate day-to-day practice while journals and books overflow with theories that claim—on paper—to be superior. The tension is epistemic, not technical: we lack a robust method for filtering durable insight from transient hype.
This chapter therefore pursues three objectives:
-
Classification. We locate supply chain on the map of knowledge—as an applied branch of economics—and show why mistaking it for mathematics, sociology, or mere record-keeping leads to systematic failure.
-
Validation. We examine what counts as evidence in a domain where randomized trials are scarce, counterfactuals dominate, and success is measured in profit rather than elegance or consensus.
-
Corruption. We analyze how actors—academics, consultants, software vendors—despite good intentions, unwittingly distort the knowledge pipeline, and outline safeguards that preserve integrity amid adversarial incentives.
Throughout this book, the singular “supply chain” denotes the discipline, whereas the plural “supply chains” refers to the concrete flows run by companies. The discipline’s role is to arm any given chain—not an idealized archetype—with decisions that raise long-run profit under uncertainty. Conversely, extending the term to whole markets drifts into general economics and blurs the practical lens adopted here.
By the end of the chapter, the reader will possess a yardstick for judging every subsequent claim—algorithmic, organizational, or technological: Does this idea allocate scarce resources more profitably when variability and adversarial incentives are accounted for? Establishing that yardstick is the prerequisite for the analyses that follow.
Economics
Supply chain must be understood as an applied branch of economics, not a free-standing discipline. Overlooking this hierarchy is perhaps the most frequent—and most severe—epistemic error in the literature. This single error explains much of why mainstream supply-chain knowledge proves barren in practice.
Regrettably, economics itself is frequently misunderstood, even though its core ideas are quite simple. The British economist Lionel Robbins gave what is typically regarded as the classic definition of economics:
Robbins’ sentence is deceptively dense; each word carries weight. Resources encompass far more than cash. Factory hours, pallet space, truck capacity, managerial attention, even goodwill with a supplier all qualify. Whatever can be employed to satisfy a want—or withheld to satisfy another—enters the economist’s ledger.
Scarcity is the universal constraint. Desires outrun means in every society, whether a medieval manor starved for arable acres, a planned economy constrained by steel quotas, or a venture-backed startup racing against its burn rate. Scarcity is not a quirk of capitalism; it is the human condition.
Alternative uses encode the bite of opportunity cost. A ton of copper rolled into power cables cannot also be stamped into cartridge cases; an hour of a planner’s focus spent chasing a late shipment cannot refine next season’s assortment. Every allocation vetoes its rivals, and that foregone benefit is the hidden price of every choice.
Because these three elements—resources, scarcity, and alternative uses—persist under feudalism, socialism, or capitalism alike, economics remains valid regardless of the social order. The mechanisms that resolve scarcity differ (prices, queues, decrees), but the underlying calculus of trade-offs does not. Supply-chain practice must therefore rest on economic bedrock, no matter how the firm is owned or how markets are organized.
Set against our definition of supply chain, the relationship becomes clear:
Supply chain is the mastery of optionality under variability in managing the flow of physical goods.
Supply chain fundamentally parallels economics in its treatment of scarcity. Here, “optionality” names the alternative uses the company must first cultivate and then exploit. Each supply chain decision utilizes a scarce resource with alternative potential. Every item bought contends with all other possible purchases, and every item manufactured competes with alternatives that could have drawn on the same materials. Similarly, each site receiving an extra dispatch unit vies with all other sites for that single item from the warehouse.
In summary, if we focus solely on allocating scarce resources, supply chain would be indistinguishable from economics. However, supply chain diverges from economics in several important ways. First, “mastery” denotes the profit-seeking stance of companies. Supply chain is an applied science; unlike the broader field of economics, it aspires to be prescriptive rather than merely descriptive. While economics explains why some individuals are wealthier than others, it does not claim to directly guide individuals toward greater wealth—except indirectly through a better understanding of market mechanisms.
Furthermore, the scope of supply chain is restricted to the flow of physical goods—a small subset of economics. For example, intellectual-property questions—patents and copyrights shaping production choices—belong to economics, except where they constrain production despite adequate physical resources.
Moreover, the scope narrows further with the element of variability; without variability, an issue would cease to be a supply chain concern. For instance, consider a warehouse process that can be nearly perfectly automated, setting a strict upper bound on pick-and-pack time—provided the arrival rate stays below a threshold. In such cases, supply chain accepts the automation as a given, without scrutinizing the economic decisions that enabled it.
Properly classifying supply chain is crucial, because the discipline is routinely approached through misguided lenses—sterile mathematics, clerical bookkeeping, or managerial platitudes—while its genuine foundation is, and always has been, economics.
Contrary to popular belief, economics is far from a “dismal science”. Granted, it cannot achieve the numerical certainty of most natural sciences; the free will of participants precludes that. That said, many economic principles can be considered as solidly established as any scientific theory can be.
Consider the well-documented case of rent controls. When a government caps rents below the market-clearing level, several predictable outcomes follow. Property owners lose the incentive to maintain or improve existing units, as they cannot recoup those costs through higher rents. Developers become reluctant to build, and long-term shortages ensue. Shortages lengthen waiting lists, reduce choice, and degrade housing quality. Such effects are observed time and again across different cities and eras, illustrating how economic principles remain remarkably consistent, regardless of cultural and temporal differences.
A nearly identical mechanism appears inside individual firms whenever a product is sold below its market-clearing price. Imagine a consumer-electronics manufacturer launching a new console at a price intentionally set 20% below the market-clearing level in the hope of “delighting the customer”. Demand instantly overwhelms the short-term production capacity. Retail channels are wiped out, waiting lists form, and—most revealingly—gray-market intermediaries (scalpers) purchase the scarce consoles at list price only to resell them online at the actual market price. The margin the company voluntarily relinquishes is captured in full by these middlemen, while end-customers still pay the higher amount. The firm forfeits revenue, frustrates loyal clients, and tarnishes its brand by appearing disorganized.
Unless the company can raise output fast enough to meet the unexpected surge, the economically sound remedy is to remove the self‑imposed price ceiling and let prices rise until demand equals capacity.7 Any other policy merely subsidizes an ecosystem of arbitrageurs whose sole contribution is to pocket the difference between the constrained price and the true clearing price.
The laws of economics are no easier to evade than the laws of electromagnetism. Failing to characterize supply chain as applied economics yields, at best, theories and practices rightly ignored by practitioners, and at worst, decisions that harm firms.
Mathematics
In academic circles, a prevalent mistake is to regard supply chain as a branch of mathematics. The book Fundamentals of Supply Chain Theory (2019) by Snyder and Shen is a canonical exponent of this misguided path. Over the past seven decades, despite over a million published papers8, academic supply chain literature has contributed little to real-world practice. Supply chains remain, with few exceptions, governed by spreadsheets. A company is more likely to profit from buying lottery tickets than from adopting any “provably optimal” inventory optimization technique. Nonetheless, thousands of researchers seem to have devoted their careers to this approach. It is astonishing that the community has persisted in such a massive error for so long.
When supply chain is treated as a branch of mathematics, it is reduced to a series of mathematical puzzles. The newsvendor model—where a newspaper vendor decides how many copies to stock under uncertain demand, with unsold copies becoming worthless—is one of the oldest and most notable examples. Similarly, the economic order quantity, defined as the order quantity that minimizes total holding and ordering costs under constant future demand, stands as another classic example. Solving these puzzles typically involves analytical methods, probability theory, or algorithms.
In this paradigm, the supply chain puzzle begins with a formal characterization of the physical flow of goods—the products, sites, demand and lead-time equations, and constraints. From this formalization, authors derive a mathematical solution—ideally with provable properties or, at minimum, a heuristic validated on (often synthetic) data.
Internal consistency is not enough; survival in contact with reality is. What matters is whether a numerical recipe survives contact with trucks, lead times, and customers. The criterion we adopt—and its consequences for this book—are stated below under Falsifiability.
Put plainly, a theory that survives solely by refusing to expose itself to reality tells us nothing about how trucks roll, how customers buy, or how pallets wait. Internal consistency guarantees correctness only inside the walls erected by the assumptions; the moment those walls meet the first promotional spike or the first customs delay, the “optimal” policy reverts to an untestable conjecture.
When treated as a branch of mathematics, supply chain puzzles remain detached from real-world conditions. For example, under a few assumptions—stationary demand, constant lead time, non-perishable goods, etc.—an inventory replenishment policy can be proved optimal. No real-world supply chain can contradict such a result: it is impossible to devise an experiment to challenge it. The solution is based solely on internal consistency derived from the problem’s assumptions.
Proponents of this approach would undoubtedly argue that the puzzles were not selected at random. On the contrary, these puzzles have been carefully crafted to accurately reflect the challenges of real-world supply chains. Moreover, many papers claim support from “real-world” datasets and report rigorous tests against them.
However, practice offers a clear and forceful rebuttal of this proposition. Companies neither use these supposedly “optimal” methods to manage their supply chains nor adopt any of their variants, not even as a source of inspiration. If these algorithms were genuinely valuable, the community would see shocks akin to those in software, where superior techniques—transactional databases, cloud computing, deep learning—are rapidly embraced.
The claim that companies are unaware of these “modern” supply chain methods does not hold up under scrutiny either. Since the late 1970s, enterprise vendors have packaged thousands of such methods. Yet, they are invariably discarded, as supply-chain practitioners swiftly revert to their spreadsheets.
Applying Occam’s razor, the simplest explanation is that these mathematically neat “puzzles” fail to produce any repeatable, profit-relevant gain once confronted with the messiness of day-to-day operations. Because they neither cut total cost nor raise service levels in a way that outweighs their implementation burden, practitioners quite rationally set them aside and return to tools—even spreadsheets—that have proven their worth.
Finally, much of the apparent “progress” in this branch of mathematics may be illusory. Research is just as vulnerable to fashion as any other branch of knowledge. In the 1960s, explicit analytical resolutions for supply chain puzzles were in vogue; in the 1990s, parametric time-series models gained popularity; and in the 2020s, Monte Carlo simulators are the trend. Authors focusing on current scientific fads do not necessarily contribute to genuine progress—or its absence—in the field. Moreover, earlier trends are labeled “outdated”, regardless of what those papers may have contributed to real-world supply chains.
Falsifiability
Having established that the mathematical posture by itself cannot steer a supply chain, we need a positive criterion for sorting knowledge that deserves to govern a pallet from what does not.
A decisive perspective on complex theories emerged in the first half of the 20^th^ century with the work of Karl Popper9. In his lifelong attempt to characterize science, Popper introduced the criterion of empirical falsification. This criterion is central to the present book. An anecdote from his life sheds light on this criterion.
In the opening chapter of his book “Conjectures and Refutations” (1963), Popper explains how his youthful experiences in Vienna around 1919 shaped his idea of falsifiability as the hallmark of science. He notes that whereas Einstein’s relativity made bold predictions—and risked being refuted by observation—Marxism (and also Freudian psychoanalysis) was perpetually “saved” by ad hoc explanations whenever contrary evidence appeared. Physicists testing Einstein’s theory welcomed, and even sought out, the possibility of falsification, using observations—e.g., starlight bending around the sun during a solar eclipse—to see if the theory held up. Physicists sought maximal exposure to reality, pursuing the angles most likely to tear a theory apart.
In contrast, Marxists would “immunize” their doctrine from refutation by reinterpreting events to fit the theory. For example, Marxist theory predicted the proletarian revolution would occur first where factory workers were most numerous—i.e., Great Britain. Yet the revolution occurred in Tsarist Russia in 1917, one of Europe’s least industrialized countries. Instead of rejecting the theory as disproved, Marxists revised the interpretation to explain why the revolution would happen first in the least industrialized country.
Obviously, physicists were the ones with the correct intellectual stance. Their intellectual inclination would end up delivering two monuments of our present understanding of the universe, namely general relativity and quantum physics. Two theories that turned out extremely capable of making predictions of empirical significance. Conversely, the Marxist theory would keep being endlessly refuted by dozens of countries that implemented reforms inspired by this doctrine, and got dismal economic results as a result, baffling their Marxist economists in the process10.
These observations led Popper to articulate falsifiability. It highlights the radical asymmetry, in science, between truth and falsehood. No theory can ever be proven true; it can only be disproved. Indeed, to prove a theory exhaustively—say, general relativity—one would have to verify that every celestial body, everywhere and always, follows the laws it states. This is obviously a wholly impossible feat. At best we can gather a long series of measurements in agreement with the theory, but beyond that, we can only extrapolate the validity of the theory to the rest of the universe. However, the converse is much more accessible: it only takes a single pair of celestial bodies11 contradicting the theory in order to prove that the theory is wrong.
Thus, at best, we can only attempt to disprove a theory. However, for such an attempt to ever take place, the theory must expose itself to an empirical validation process that could, in principle, fail. Popper referred to such theories as falsifiable. It follows that the longer a theory survives relentless experimental attack, the more confidence it earns. In contrast, an unfalsifiable theory cannot be disproved; such theories do not belong to scientific knowledge. Falsifiability is now widely recognized as one of the central tenets of science.
Circling back to supply chain, falsifiability reveals what is wrong with the purely mathematical posture: it is immune to refutation. It belongs to applied mathematics, not to the natural sciences. As a result, the correctness of the numerical recipes derived from this perspective is shallow. It merely reflects consistency with an initial bundle of assumptions. Those recipes are left entirely vulnerable to the correctness of their premises, but also to even more mundane issues.
Falsifiability matters because supply chain is judged in profit, not in elegance. A recommendation that cannot be shown false when shelves stay empty or when inventory bloats is not operational science—it is mathematical storytelling. Mathematical rigor is indispensable, provided the premises invite empirical challenge.
The incentives of academia
Before examining concrete illustrations of the oversized assumptions that cripple these models, we must ask why academia keeps championing them. The answer lies in the incentives that shape scholarly careers: journals reward novelty and formal elegance—not operational impact—so research naturally gravitates toward puzzles that are publishable rather than practicable.
In The Future of Operational Research is Past (1979), Russell Ackoff identifies a series of issues undermining his field. In contemporary terms, “operational research” is synonymous with “supply chain”, and Ackoff’s critiques remain entirely relevant today. Those problems include:
-
Solving made-up problems;
-
Taking the objective for granted;
-
Ignoring system properties;
-
Assuming the future is knowable;
-
Being inward-looking, and increasingly so;
-
Rationalism passing for reason.
More than four decades later, it may seem astonishing that academia still overlooks those foundational issues, especially given Ackoff’s significant reputation. Yet this behavior becomes less surprising when one considers the incentives.
Academics are primarily assessed on two tightly coupled axes of evaluation. First comes the volume of papers produced and the impact those papers achieve inside the scholarly echo chamber—impact here meaning little more than how often other papers cite them. Whether the cited idea ever trims a procurement bill, shortens a lead time, or prevents a stockout does not register in this metric.
The second axis is teaching performance, which in practice reduces to the lecturer’s aptitude for transmitting that same citation-oriented material to the next cohort of would-be scholars. This twin yardstick creates a self-referential loop. Intellectuals are judged solely by how well they persuade other intellectuals; classrooms are staffed and syllabi written by those who have already mastered the art of persuading peers. Empirical effectiveness—the capacity of an idea to survive contact with forklifts, tariffs, or customers—never enters the equation.
Consequently, mathematical problems fitting the supply-chain theme offer an inexhaustible stream of publishable statements and resolutions. Adding constraints, modifying demand equations, altering network topology, or incorporating additional terms into objective functions produces nearly endless variants. Each variant serves as ideal material for an academic publication. Moreover, the specific details from a business willing to share even modest datasets can provide a unique—and thus novel—perspective on the subject. This inevitably fosters the illusion that resolving an isolated puzzle holds significance for the domain at large.
Thus, academia relentlessly treats supply chain as a branch of mathematics because it is the most expedient way, in terms of time invested, to get papers published. Furthermore, these papers exhibit the fashionable hallmarks of “great science”—such as equations, algorithms, and quantitative results. By categorizing supply chain research as applied mathematics, authors avoid the rigorous falsification challenges that natural sciences routinely confront.
In classroom teaching, because promotion hinges on student evaluations and pass rates, instructors have every incentive to present material that mirrors what their colleagues will set in examinations—again privileging continuity over confrontation with reality.
Also, mathematics is unusually easy to teach—assuming the speaker understands the material. Unlike in the natural sciences, no dubious compromises are necessary to fit the presentation into the allotted time. The speaker can simply tune the number and complexity of puzzles to fit the clock.
Furthermore, when it comes to grading, mathematical puzzles are maximally convenient. Their number and complexity can be precisely adjusted to fit the exam context, enabling nearly flawless, objective grading with minimal effort. Unfortunately, this results in puzzles that bear little resemblance to the messy, ambiguous, and complex realities of actual supply chains. Intellectuals teach future intellectuals the material authored by past intellectuals, with no external milestone to break the circle.
No matter how poorly this approach—treating supply chain as applied mathematics—performs in practice, it still addresses the enduring concerns of academic career seekers. Thus, as Ackoff illustrates, merely proving that this research approach is misguided is insufficient; the underlying detrimental incentives within academia must be addressed—a task entirely beyond the scope of this book.
Let’s clarify that there is nothing inherently wrong with using the supply chain theme in mathematics—or in fields such as analysis, algorithms, and statistics—if a mathematical problem is best illustrated by a supply chain metaphor. For instance, consider the famous Seven Bridges of Königsberg problem, solved by Leonhard Euler in 1736. The challenge was to devise a route through the city that crossed each bridge exactly once. Euler proved that no solution exists, introduced the concept of graphs, and thereby paved the way for the field of topology.
Under these circumstances, a paper is worthy only if it possesses intrinsic mathematical merit: Does it reveal a hidden truth or structure? Is the exposition particularly elegant? Does the method pave the way to solve other, seemingly unrelated, mathematical problems? Otherwise, the work is deemed worthless. Simply using a supply chain theme cannot confer merit on work that fundamentally belongs to the realm of mathematics.
Any reader familiar with the literature will know that supply chain papers meeting this standard are exceedingly rare.
The pitfalls of safety stocks
Demonstrating the shortcomings of the vast literature that treats supply chain as mathematics is daunting; with roughly a million papers to challenge, the task borders on insurmountable—a vivid illustration of Brandolini’s law, which notes that refuting nonsense usually demands an order of magnitude more effort than producing it.
To keep the discussion manageable, we focus on one widely accepted concept—safety stocks—to illustrate the general shortcomings of approaching supply chain as mathematics.
Safety stock is the buffer required to achieve a desired service level for a given SKU (stock-keeping unit). Here, the service level is defined as the frequency with which a stockout is avoided during a given period. The model assumes demand during lead time is normally distributed and lead time is constant. These assumptions yield a closed-form formula for the required stock.
Since the calculation is of historical interest only, it is omitted here. Suffice it to say, the model yields an explicit formula that can be easily computed in a spreadsheet with minimal effort from both the programmer and the computer.
Yet safety stocks are fundamentally flawed—“hazardous stocks” would be apter—because the mischaracterization yields poor outcomes. As a silver lining, safety stocks serve as a simple litmus test to identify gross incompetence in the field of supply chain.
Safety stocks present three major issues, each of which is sufficient to preclude satisfactory outcomes when applied in a real-world supply chain. The first two are categorical errors, while the third is technical.
The first—and most damaging—flaw is methodological isolationism. Safety-stock formulas examine every SKU as if it were the only claimant on the balance sheet, striving to compute an “optimal” quantity for that item in a vacuum. Yet each extra unit ties up cash that could have funded stock for another product or been deployed to any more profitable use. Indeed, all SKUs compete for the same finite pool of working capital, and any model that ignores this rivalry is blind to the very trade-offs it is supposed to orchestrate.
Outside micro-companies, no supply chain deals with a single SKU; firms typically manage hundreds, if not thousands. Moreover, replenishing one SKU is fundamentally different from managing replenishment across many SKUs. The two problems are radically distinct. Safety stocks fail even to acknowledge that SKUs cannot be treated in isolation; their very definition sidesteps the problem that must be solved: the allocation of money for inventory.
Second, even if we momentarily accept treating each SKU in isolation, the safety-stock recipe remains hopelessly incomplete. It hinges on picking a “target service level” — a percentage plucked from thin air — and then treating this number as if it possessed a built-in link to profit. The formula never explains why 95% should outperform 92% or 97%; it simply delegates the most difficult judgment call to managerial fiat and proceeds as if the matter were settled.
Proponents sometimes reply that a higher service level will “generally” boost profitability, but correlation is not causation. Set the level too high and dead stock accumulates the moment demand shifts; set it too low and frustrated customers defect to competitors. Because the mathematics permits any target between 0% and 100%, the practitioner ends where they began: facing the real economic question of how much capital to commit to which items. The safety-stock apparatus does nothing to answer it.
The third—and final—flaw is statistical naïveté. Real-world demand is almost never bell-shaped: it is sparse for long stretches, then jumps without warning, and its extremes occur far more often than a normal law would allow. Lead times show similar unruliness, clustering around distinct modes—“on-time”, “customs delay”, “supplier stockout”, and so on—rather than hugging a single, symmetric peak. These patterns are the norm, not outliers, and any model that ignores them starts from the wrong coordinates.
Textbooks and software cling to the normal curve because it keeps the algebra tidy, not because it matches the data. In principle one might swap in zero-inflated, heavy-tailed, or mixture distributions to repair the theory, but doing so strips away its lone virtue: simplicity. Once the arithmetic becomes opaque, the safety-stock recipe loses its only practical appeal while still inheriting all the earlier categorical mistakes. If a method is neither realistic nor simple, nothing remains to recommend it.
The vacuum left by mathematically sterile academia is, alas, quickly filled. Into it step consultants and business-book authors who swing the pendulum to the opposite extreme: abandoning numbers for narratives. Where the academic fallacy treats supply chain as a puzzle bereft of incentives, the consultant fallacy treats it as a story bereft of quantification. Both camps ignore the economic core, merely from different ends of the spectrum. The next section examines how this sociological turn, though wrapped in inspirational prose, proves no more reliable than the equations it claims to supplement.
Sociology
If the academic path sterilizes supply chain into abstract puzzles, the consulting path swings to the opposite extreme: it recasts the discipline as a catalog of sociological observations and leadership bromides. Business consultants have therefore dominated supply-chain publishing for the last four decades, framing every problem in terms of culture, teamwork, and “best practices.” Their output is vast—well over 10,000 titles appear under “Supply Chain Management”12. Dynamic Supply Chains (2015) by John Gattorna is a canonical example of the genre—and of its epistemic weakness: persuasive anecdotes replace testable evidence, leaving the reader with stories rather than substantiated knowledge.
This literature belongs squarely to the self-help genre. Its intended readership is the corporate executive—or the manager who hopes to become one—who believes that mastering a new buzzword will both lift company performance and accelerate personal promotion. The consultant-author, for his part, adopts the posture of a guru whose insight can be rented by the day.
Each book then unfolds along two predictable axes. The strategy axis offers a handful of slogans—typically three to five—that promise to explain how one should “think” about the supply chain; the discipline is sliced into neat quadrants, waves, or maturity levels. The organization axis converts those slogans into a checklist of rituals: steering committees, cross-functional workshops, responsibility assignment matrices, and the like, all presented as turnkey pathways to execution.
To sustain momentum, the narrative is laced with borrowed authority: sound bites from Fortune 500 executives, quotations from Navy SEALs about resilience, and blow-by-blow retellings of last-minute sports victories. Besides briefly enlivening the prose, a sports anecdote—or an athlete’s quote—adds no evidential weight to supply-chain arguments. Celebrity in athletics or entertainment has no bearing on the allocation of pallets and purchase orders. Such cameos create the illusion of battle-testing; in fact they deliver color, not proof.
The genre relies on a familiar first trick: buzzword inflation. Terms such as “agile”, “digital twin”, or “AI-driven” are repeated until they sound inevitable, yet their definitions remain elastic enough to fit any circumstance. Behind the shimmer, one finds only evergreen truisms—“the world is changing, you must adapt”—rebranded as revelations. Taxonomies that slice firms into “leaders” and “laggards,” or “innovators” and “followers”, cloak tautologies in flow-chart form and create the illusion of depth where none exists.
The second trick is manufactured urgency. Authors spotlight a handful of allegedly neglected levers—supplier intimacy, customer centricity, sustainability, culture—and intone that immediate escalation to the C-suite is vital. This “act-now-or-perish” framing is impossible to falsify in a boardroom setting. It deftly diverts attention (and budget) toward the consultant’s favored initiative without ever weighing the opportunity cost imposed on all the other constraints a functioning supply chain must honor.
The final trick is borrowed authority. Case studies name-drop admired brands, parade selective success metrics, and sprinkle quotations from athletes or Special-Forces veterans. The implied syllogism is specious: “Company X is famous; Company X endorses this framework; therefore, the framework creates fame.” Correlation is quietly traded for causation, and the failures that would break the spell are screened out of view.
These rhetorical devices are not harmless entertainment; they crowd out the quantitative reasoning that genuine improvement requires. By wrapping platitudes in prestige and urgency, the literature flatters executives into feeling decisive while leaving the actual flow of goods untouched. Strip away the logos and the motivational anecdotes, and nothing remains that can survive analytical scrutiny.
In other words, the consulting canon treats supply-chain dysfunction as a human-relations problem whose resolution lies in reshaping group behavior. Its favored variables are roles, rituals, and narratives; its instruments are workshops, charters, and incentive grids. This viewpoint is textbook organizational sociology: it studies how collectives maintain cohesion, how status is negotiated, and how norms are enforced, then prescribes ceremonies meant to realign those social constructs. Pallets, containers, and lead times appear only as props in a narrative whose real protagonists are departments vying for influence. Whether tonnage actually flows more smoothly is deemed an emergent by-product of better social harmony. This literature is therefore sociological in method and aim—even if it rarely advertises itself as such.
Even if this literature contained genuine insights—which it seldom does—treating supply chain chiefly as a sociological endeavor leads improvement efforts astray. Consider a thought experiment: imagine a near-future platform in which a single autonomous artificial intelligence (AI) ingests every demand signal, lead-time update, and cost curve, then issues—without human intervention—every purchase order, price change, and routing instruction. In that world there is nothing left to “align” or “empower”; the entire scaffolding of committees, workshops, and incentive grids evaporates overnight. Yet the economic substance of the discipline endures: scarce resources must still be allocated under variability. The AI would lean on principles of optionality, opportunity cost, and risk—core economics—not on any theory of inter-departmental politics. The thought experiment clarifies the hierarchy: organizational constructs are temporary workarounds we erect only because people, slower and scarcer than CPUs, must coordinate. Automation is already eroding this sociological wrapper—pricing bots, routing optimizers, and allocation engines now perform tasks that once filled entire departments—just as a laptop replaced the actuarial tables that formerly justified platoons of clerks. Until full automation arrives, humans will stay in the loop. The compass that guides them must remain economic and operational, not sociological.
Granted, some supply chain challenges still resist software. Once the necessary steps are explicit, it is acceptable to rely on human intelligence as a stopgap until state-of-the-art methods can automate them. Where workload exceeds one person’s capacity, add people and organize them accordingly; even then, sociology should be a last resort. Indeed, to insist that an intellectual task “requires human intelligence” is to admit: “I have no clear idea how to solve this problem, but if push comes to shove, someone will figure something out.”
For more than two centuries, it has been a recurring mistake to treat human involvement as a virtue in itself. The Luddites—British weavers and textile workers who opposed mechanized looms and knitting frames in the early 19^th^ century—are the canonical example of this mistake. In supply chains, human involvement in the daily micromanagement of millions of SKUs is not inherently desirable. The absence of mechanization imposes predictable limits—long processing delays, reduced reactivity, and mediocre allocations—because operators cannot weigh all relevant quantitative factors.
Taken together, these consultant-driven tricks explain why glossy frameworks so often leave containers, prices, and purchase orders exactly where they started. They turn resource allocation into group therapy, mistaking harmony for profit. To see the hard cost of confusing sociology with supply-chain mastery, we now examine the flagship doctrine of this literature—Sales and Operations Planning.
The pitfalls of S&OP
Sales and Operations Planning (S&OP) has been heralded for four decades as the cure for functional silos, yet it epitomizes the sociological fallacy just exposed. Its elaborate calendar of meetings seeks consensus around a single—highly uncertain—forecast and then asks every department to pledge allegiance. Renamed variants such as integrated business planning (IBP) inherit the same weakness: they treat pricing, purchasing, and capacity as a corporate ritual rather than an economic optimization. Because S&OP is both influential and symptomatic, it will serve as the lone exhibit in the critique that follows, though the argument applies equally to all its look-alikes.
Since the rise of the giant corporation in the 19^th^ century, specialization has supplied most of what we label “economies of scale”. By narrowing their focus, teams deepen expertise, lift productivity, and make outcomes more predictable. Yet every new specialty opens another seam that must be stitched back to the whole, and the stitching grows combinatorially harder as the organizational tapestry expands.
When coordination lags, the misfires are dramatic. Marketing can stoke demand that plants cannot meet, squandering advertising budgets and disappointing customers. Manufacturing can flood warehouses with units the market will not absorb, forcing fire-sale discounts or outright write-offs once carrying costs eclipse any plausible margin.
Sales and Operations Planning (S&OP) was devised in the early 1980s as a ritualized cease-fire between the factions just described. Its promise was simple: convene every function around a single calendar checkpoint, agree on one forward view of demand, and cascade that view into synchronized supply, capacity, and financial plans. In theory, the meeting guarantees that the people who create demand (sales, marketing, merchandising) and the people who fulfill it (procurement, manufacturing, logistics) leave the room committed to the same numbers and the same constraints.
The framework gained momentum in the 2000s when ERP vendors and consultancies began selling S&OP dashboards, maturity models, and benchmarking surveys. Despite the glossy tooling, the mechanics remain spartan: a rolling 18- to 24-month horizon, a consensus forecast sometimes styled “one set of numbers”, and a monthly sequence of pre-reads and workshops that massage those numbers until every vice president can sign off.
In historical perspective, S&OP is less a breakthrough than a modern sequel to Alfred Sloan’s executive production conferences at General Motors in the 1920s. The vocabulary has changed, but the intent is identical: impose an institutional routine that keeps functional empires from drifting apart when variability threatens to pry them loose.
Despite its longevity, S&OP collapses under close inspection. Four structural defects undermine the practice and show why, from an epistemic standpoint, it cannot be trusted to steer scarce resources.
-
The forecast illusion. S&OP begins by assuming that a single, time-series forecast of demand not only exists but is precise enough to govern capacity, pricing, and inventory months in advance. That premise treats an uncertain future as a known constant, turning the hardest part of the problem into a spreadsheet input. By disguising uncertainty as certainty, the ritual severs the feedback loop that would otherwise test whether those numbers ever deserved belief.
-
The decision vacuum. Having conjured a consensus demand forecast, S&OP then hands off every quantitative decision—procurement lots, production schedules, routing plans—to whatever heuristics each functional silo already uses. The framework contributes no economic criterion for choosing one allocation over another; it merely notarizes choices made elsewhere. Epistemically, this is bookkeeping masquerading as optimization.
-
Ritual overreach. The monthly cadence of pre-meetings, alignment workshops, and executive reviews is presented as the cure for miscoordination, yet the cure scales with headcount, not with the physics of the flow. Every extra SKU, lane, or promotion demands another slide deck, not a sharper algorithm. The complexity is accidental—created by meetings—rather than essential—created by atoms and lead times.
-
Arbitrary taxonomies. S&OP partitions time into operational, tactical, and strategic buckets and reduces reality to a two-axis “demand vs. supply” matrix. These partitions are rhetorical conveniences, not empirical discoveries; they can be rearranged at will without changing economic outcomes. A theory whose parameters are free to float cannot be falsified and therefore cannot accumulate reliable knowledge.
Together these defects explain why companies that adopt S&OP often return to homegrown spreadsheets: the practice offers process theater without predictive power. A meeting can enforce temporary consensus, but it cannot reveal which decision maximizes long-run profit under variability and adversarial incentives. Until a framework exposes its assumptions to empirical challenge and grounds its prescriptions in opportunity cost, it belongs to organizational folklore, not to the science of supply chain.
We now turn to economic history to examine how similar epistemic missteps have played out in the past—and what lessons they hold for the present.
Economic history
Economic history catalogs what actually happened—the prices paid, the goods moved, the crises endured. It is indispensable for context, yet, as Ludwig von Mises stressed in Human Action (1949), facts do not interpret themselves. Only a prior economic theory can assign meaning by tracing cause, effect, and counterfactuals. In short, economics must precede economic history; we first need a logical framework before we can usefully retell the past.
The same ordering governs supply chain. Dashboards brimming with lead-time histograms, service-level charts, and inventory turns are nothing more than chronicles. Without a theory of opportunity cost and optionality, an analyst cannot tell whether a 20% stock reduction was brilliance or negligence. Data become knowledge only when filtered through an economic lens; otherwise, the exercise is bookkeeping masquerading as insight.
Yet in day-to-day practice, the distinction between chronicle and explanation is routinely blurred. Analysts fill dashboards, vendors publish benchmark reports, and consultants parade before-and-after case studies—all of which recount what did happen while remaining silent about what could have happened under different choices. The record tells a story; it never diagnoses the mechanics that generated that story.
Consider a retailer that advertises a 20% inventory reduction. Absent a theory of opportunity cost, the headline is meaningless: was working capital truly released with no loss of service, or were shelf-outs simply postponed to a quarter yet to be reported? The ledger contains no rows for the sales that never materialized, the assortment expansions never attempted, or the price experiments never run. History records only the path taken, never the adjacent ones left unexplored.
Benchmarks magnify the problem. Ranking companies by “days of supply” or “inventory turns” invites imitation yet ignores strategic asymmetries such as assortment breadth, service promises, or pricing power. Copying the metric without reproducing the context is cargo-cult optimization: the ritual is preserved, the economics are not.
The lesson echoes the critiques leveled at mathematics and sociology: economic history is an auxiliary source of insight, not a foundation. Facts collected in arrears are indispensable, but they must be filtered through a forward-looking economic lens that weighs counterfactuals and opportunity costs. Without that lens, history becomes numerate storytelling—comforting to read, powerless to steer the next container. Mastery therefore starts with theory, tests that theory against experience, and only then lets the ledger guide future allocations.
Auxiliary sciences
Auxiliary sciences are the disciplines that supply chain routinely relies on for facts, proofs, or tools—statistics for uncertainty, computer science for automation, empirical psychology for human bias, aeronautical engineering for fuel efficiency, and so on. The phrase comes from historiography, where since the sixteenth century fields such as paleography or dendrochronology have been called auxiliary because they supply evidence without writing history. The hierarchy is epistemic, not pejorative: supply chain is subordinate to whatever durable truths those sciences establish—if statisticians refine the meaning of a confidence interval or aircraft engineers demonstrate a cheaper flight profile, the practitioner must adapt at once. What matters is that the knowledge is reliable, not the route by which it was obtained.
Subordination, however, does not imply fusion. Supply chain is not a multidisciplinary stew of mathematics, software engineering, psychology, and logistics. It remains a distinct applied branch of economics whose sole mandate is to allocate scarce resources under variability. Auxiliary sciences are consulted only insofar as their findings help to fulfill that mandate; their internal debates, jargon, and prestige contests can be safely ignored. A practitioner needs just enough familiarity to choose the right tool and to notice when it fails—no more, no less.
The required familiarity depends on packaging. When an auxiliary result is embedded in a robust artifact—a barcode scanner, an optimization engine, an integrated circuit—the supply-chain expert can treat it as a black box and concentrate on the economic trade-offs. When the packaging is thin—as is still the case with most behavioral-economics research—the expert must open the box, grasp the limits, and compensate manually.
Historians offer the analogy. Dendrochronology can date a shield to the summer of 1519 and thereby refute any claim that it was wielded at the 1515 battle of Marignano, yet no historian confuses the study of tree rings with the writing of history. Supply chain should cultivate the same discipline: exploit every auxiliary science ruthlessly, but never surrender the steering wheel.
The subordination principle is clear when microbiology sets non-negotiable boundaries. Laboratory work shows that most food-borne pathogens enter exponential growth once product temperature rises above 5 °C for more than a few hours. Cold-chain guidelines therefore mandate that chilled meat leave the plant below 4 °C and never exceed 7 °C before it reaches the customer. Whether planners can recite the growth curves of Listeria is irrelevant; once microbiologists fix these thresholds, the supply-chain expert must decide whether to pay for reefer containers, run denser delivery routes, or shorten the distribution radius—but the temperature constraint itself is not up for debate.
Materials science offers the same lesson. Compression tests reveal that a standard 7 mm double-wall corrugated carton collapses once the top load exceeds roughly 8 kN. A distribution center that stacks such cartons six pallets high has therefore capped its per-carton weight at about 18 kg—unless it upgrades to triple-wall packaging, lowers the stack height, or invests in racking. The auxiliary science supplies the boundary conditions; supply chain weighs the economic trade-offs that live inside those boundaries.
Supply chain defers to its auxiliary sciences because those fields possess faster and sharper ways to separate truth from error. In physics or chemistry a benchtop experiment, run before lunch, can annihilate a bad hypothesis at negligible cost; results are quick, repeatable, and decisive. Supply-chain claims, like those in economics, can be vetted only in the messy world of plants, trucks, and stores. Field trials stretch over months, burn real money, and rarely deliver a crisp yes-or-no verdict. Because its feedback loop is slower and noisier, supply chain must borrow the rigor of its auxiliaries while accepting that its own knowledge will accrue more gradually.
Nonetheless, supply chain does not inherit the objectives of its auxiliary sciences; it merely buys their findings as raw inputs. Aircraft engineers may prove that inserting a mid-route refueling stop trims fuel burn by eight percent on an ultra-long-haul flight, yet the practitioner must still weigh that saving against longer transit time, extra landing fees, crew overtime, and possible customer churn. The criterion is economic payoff, not technical elegance: every resource—jet fuel, hours, capital, goodwill—carries an opportunity cost that must be priced before a decision is made. Auxiliary sciences deliver facts; supply chain assigns value.
Supply chain may appear multidisciplinary because its daily practice borrows statistics, computer science, mechanical engineering, and even behavioral psychology. This impression confuses tools with objectives. As a science, supply chain is a branch of applied economics whose single mandate is to decide how scarce resources should flow through space and time. Auxiliary sciences merely furnish measurements, algorithms, or machines that serve that mandate; they do not redefine it. By contrast, the natural sciences bleed into one another precisely because atoms know nothing of the departmental walls that separate physics from chemistry or biology—their borders are pedagogical conveniences. Supply chain is therefore not multidisciplinary in the epistemic sense—only in the practical sense that it consumes many imported results.
Whether a practitioner must dive into an auxiliary field depends on two levers. Economic leverage measures how much profit hinges on the ancillary knowledge; packaging quality measures how much of that knowledge comes embedded in reliable artifacts. The higher the leverage and the poorer the packaging, the more personal mastery is required.
Economic leverage is the profit sensitivity tied to an auxiliary science. If tweaking a clean-room’s airflow rate saves only a token amount per batch, leverage is low and the setting can remain a fixed procedure. By contrast, shortening the residence time of a continuous chemical reactor by just five minutes can lift hourly throughput by six percent, yet it also depresses conversion yield and shifts the slate of co-products: the primary grade sells at a premium, while the secondary and tertiary streams trade at a discount or must be reprocessed. The extra volume, the downgraded yield, and the altered co-product mix ripple through inventories, contract allocations, and price tiers. Because the trade-off is large and the control logic hides inside dense process diagrams, planners must understand the reaction kinetics well enough to weigh capacity against margin before authorizing the change. Deep expertise is therefore required only where high leverage meets thin packaging; elsewhere the auxiliary field may safely remain a sealed module.
Conversely, integrated circuits illustrate superb packaging. They embody a century of quantum mechanics behind a five-volt interface; planners exploit the computational power without ever studying band theory. Jet engines hide thermodynamics and materials science just as neatly; a fuel-burn table is all a network optimizer needs. At the opposite end lies empirical psychology. Findings such as anchoring, loss aversion, and hyperbolic discounting reach managers mostly as prose. Because the packaging is thin while the leverage—pricing, promotion, negotiation—is high, supply-chain leaders must grasp the concepts themselves or risk repeating the biases the field documents.
A practical rule of thumb follows: deep expertise is needed only where high leverage meets thin packaging. When one of those levers is absent, the auxiliary science can stay a sealed module.
In short, auxiliary sciences push the frontiers while supply-chain practice decides when those frontiers matter. Whenever statistics yields a tighter interval, materials science a lighter pallet, or computer science a faster solver, the economic leverage of the corresponding decision shifts—and so must managerial attention. Conversely, once a breakthrough is sealed inside a reliable artifact, it can be treated as infrastructure and mentally offloaded. The craft is to keep a live map of which pieces still demand scrutiny and which have crossed the “black-box” threshold. Neglect that map and a vacuum appears, soon filled by exaggerated vendor pitches and consultant folklore—a pattern examined next under the heading of epistemic corruption.
Epistemic corruption
When a knowledge system importantly loses integrity, ceasing to provide the kinds of trusted knowledge expected of it, we can label this epistemic corruption. Epistemic corruption often occurs because the system has been co-opted for interests at odds with some of the central goals thought to lie behind it. Epistemic Corruption, the Pharmaceutical Industry, and the Body of Medical Science (2021), Sergio Sismondo.
Much of what passes for supply chain knowledge today is corrupt—it fails to improve supply chains even as it claims to do so. Yet no grand conspiracies, evil overlords, or bribes are at work. Authors who produce severely biased supply chain material are rarely dishonest—more often than not, they are unaware of the problem. The bulk of the epistemic corruption plaguing supply chain stems from structural perverse incentives. The remainder of this chapter addresses enduring misconceptions whose damage extends far beyond supply chain.
As an applied branch of economics, supply chain knowledge promises wealth—making companies richer, sometimes wealthy. Every supply chain practitioner or in-house team is driven to develop superior knowledge to outcompete rivals. Consequently, they have little incentive to publicize their discoveries; if they uncover a counterintuitive yet effective heuristic for inventory optimization, companies prefer to protect it as a trade secret.
The most vocal purveyors of supply chain “truths” are those who monetize advice—academics, consultants, and software vendors—each operating under incentives that skew what they choose to publish.
First, academia: funded by private or public sources, professors vie for students and recognition through publications—a classic case of “publish or perish”. Academic authors seldom have the opportunity to verify their supply chain claims in practice; even when they engage with companies, their careers remain largely insulated from real-world outcomes. In fact, the tendency to view supply chain as a collection of mathematical puzzles underscores the influence of structural incentives.
Imagining and solving mathematical puzzles requires limited ingenuity, especially when both the new puzzle and its resolution are mere variants of known instances. Since the validity of the resolution is purely a matter of internal consistency (as in any mathematical proof), the author is relieved of the burden of seeking real-world validation. Using real-world data in synthetic benchmarks is little more than a smokescreen; it still does not show what outcomes deployment would have produced.
Second, consultants, whose clients are corporate managers. Unlike academics, consultants face the real-world consequences of their theories, and this liability often becomes an asset. When a manager doubts a project, a consultant is brought in as a buffer: success earns the manager credit, while failure falls on the consultant. Aware of this dynamic, managers rarely dismiss consultants with mediocre records. More broadly, consultants typically avoid offering firm recommendations, as their role is fundamentally to never say “no” to a client.
Third, software vendors, including the present author. Engineering enterprise software is a slow and messy process. The final product is usually opaque and convoluted—were it otherwise, it would be marketed as a self-service product for individuals and small businesses. The clients of software vendors are also managers; however, the managers who select the vendors are typically not the ones who use the product. Managers often consult subordinates to make a more informed decision, and it is here that adversarial incentives enter. Most software products are intended, at least in part, to improve productivity, which often involves reducing headcount. Relying on subordinate opinions systematically favors the vendor that best preserves the status quo (and thus headcount). Since managers often lack the time to acquire the expertise to assess a product, they defer much of the evaluation to third-party authorities.
Because of this dynamic, the case study has become the canonical promotional instrument in enterprise software. Since so much of the field’s “evidence” now flows through this channel, it warrants close examination in the next section.
Case studies
Of the channels through which these adversarial incentives distort evidence, one dominates: the corporate case study. Case studies are elaborate infomercials intended for a corporate audience—nothing more, nothing less. Yet many companies operate under the dangerous delusion that case studies are truthful, especially when the cited client is reachable; the reality is otherwise. At best, a case study clarifies what the vendor is trying to sell; it cannot assess how much the quoted client benefited, let alone what the vendor would deliver to another company. Let us dispel this delusion.
A “software” case study suffers from two intrinsic, independently fatal flaws: perverse incentives and selection bias.
There is no doubt the vendor is heavily incentivized to embellish—magnifying benefits and suppressing issues that undermine returns. The client’s voice is meant to provide a reality check; in practice, the client is just as incentivized as the vendor to embellish. Indeed, an enterprise sale requires one or more executives to champion the option. Those champions are effectively vouching for the vendor13. Moreover, their claims will be backed by highly convincing metrics. Supply chains are complex enough to guarantee that, no matter how dysfunctional the initiative, a shortlist of favorable metrics can be identified. If needed, metrics can even be quite blatantly forged. No one—outsiders least of all—except those with intimate knowledge of the initiative will be in a position to contradict the case study’s claims. Corporate case studies are, by design, not reproducible.
In the presence of such an egregious conflict of interest magnified by a complete lack of accountability, bias is inevitable. Even individuals of sound character, known for their honesty in daily affairs, can succumb to distorted incentives when reputations or livelihoods are on the line. This is not to impugn their moral fiber, but rather to recognize the powerful sway that professional pressures, peer perception, and personal ambition can exert on otherwise admirable people. When a project’s success translates into career advancement, speaking invitations, or intra-company prestige, even the most ethical may find it nearly impossible to be neutral in their assessments. The conflict of interest is simply baked in.
For those reasons, calling the client to “verify” a case study is entirely ineffective. Barring the rare case of outright fabrication, the managers involved will gladly confirm everything that was written—exactly as their incentives predict. Worse, the “human touch” lends undue credibility to the claims, simply because most corporate managers make a favorable impression on whoever reaches out to them14.
Even granting, implausibly, that the client is truthful, the case study—as a method—is still terminally flawed by selection bias. Vendors serve many clients, and case studies are not picked at random. The vendor naturally picks the most favorable cases. This vendor-driven selection guarantees massive bias.
Indeed, a vendor may be incompetent—most initiatives quietly folded as failures—yet still display successes. All it takes to achieve the optics of success in enterprise software is to take over a mess left by another incompetent vendor.
Incompetent vendors abound, and companies that pick them demonstrate a lack of judgment. It is therefore unsurprising that a company that once picked an incompetent vendor ends up picking another similarly incompetent vendor once the first failure is discreetly acknowledged. Yet, for this second round, a few lessons will have been learned, and some severe mistakes will be avoided. Thus the second vendor, despite being just as incompetent as the first, will most likely deliver something comparatively better than the first. Even the crippled can clear the bar if it is set sufficiently low. Out of this relative success, a case study glorifying the second vendor’s competence will follow.
It is futile to lament the biases of authors in the field of supply chain, just as it is futile to hope that one day an “unbiased” author will appear. There is no such thing as a neutral observer; putting any supply chain theory to the test requires the cooperation of a profit-driven company. Remaining neutral is tantamount to remaining incompetent. There is no solution to be found in the ivory tower of pure research. Adversarial incentives, and the biases they produce, should be considered an intrinsic feature of supply chain.
This means the methods used to produce supply chain knowledge should be assessed for their capacity to mitigate, rather than amplify, those biases. Ideally, those methods should be designed and directed to this very purpose.
Negative knowledge
Having surveyed how supply chain “knowledge” goes wrong, we need a practical countermeasure. The most reliable first step is negative knowledge: a curated register of what predictably fails, kept close at hand and enforced as a veto against attractive but ruinous schemes. The brief that follows explains why this register matters and how to use it.
If positive knowledge steers the mind in the right directions, negative knowledge steers it away from the wrong ones. Negative knowledge is largely absent from modern education systems. Teaching is seen as an exercise in imparting “valid” knowledge to students. Even within higher education circles, lectures where professors venture into topics such as “widespread delusions” or “what sounds good but doesn’t work” are few and far between. As a result, negative knowledge is largely absent from the supply chain literature, and supply chain is unremarkable in this regard.
Supply chain is, however, remarkably suited for negative knowledge. Failed supply chain initiatives are plentiful, yet too often brushed aside when they should be treated as a precious source of negative knowledge. Any company that has run a sizable supply chain since the 1980s and still operates on spreadsheets rather than genuinely automated systems has almost certainly endured half a dozen such fiascos. The irony is as old as the late 1970s, when software vendors first pledged complete automation, raising expectations they have yet to fulfill. Instead of letting these mishaps fade into memory, a company should examine each one in detail, producing post-mortems that upper management safeguards like crown jewels. By dissecting what went wrong—whether it was an inflated sales pitch, a muddled chain of command, or a stubborn legacy process—the organization learns to spot early signs of trouble and avoid repeating the same pattern a few years later.
A firm might, for example, embark on a grand plan to implement a fully automated inventory system across dozens of warehouses. The initiative can fail for any number of reasons: newly introduced software contradicting decades-old procedures, conflicting demands from different departments, or a vendor’s technology proving inadequate once the shiny demos give way to real-world conditions. Without a proper post-mortem, the firm risks sleepwalking into another, eerily similar venture three years later. A carefully documented failure, however painful to revisit, strengthens the company by exposing the pitfalls that slick promises or sloppy leadership create. Those who do not learn from their failed initiatives are doomed to bankroll the same mistakes again—and in supply chain, the costs only compound as the years go by.
When a supply chain fiasco comes to light, it deserves special attention and greater credence precisely because no one has an incentive to expose it. The parties involved almost always keep their failures hidden, so when a calamity becomes public, it is usually too big or too costly to be swept under the rug. This negative knowledge should be safeguarded and studied, as details of a failure do not spread easily and are nearly impossible to find in books or academic papers. A manager hoping to avoid the same expensive lesson should seize these rare disclosures and examine them carefully, for each one reveals not just a bad outcome but the bad idea that led to it.
Negative knowledge endures because its foundations do not rest on passing quirks of the market. When a method or a paradigm is defective at its core, no change in circumstances will ever fix it. Supply chain abounds in such misfires. Stack ranking15, for one, might seem straightforward; but in a field that demands close cooperation among teams, it corrodes the very trust those teams need to function. Likewise, imposing a normal distribution to model demand may look neat on paper, but supply chain data rarely follow a tidy bell curve. The demand shocks and rare events that shape real-world operations are exactly what this model overlooks. These flaws are not products of the moment, and it is wholly unreasonable to imagine that, with enough tweaking, stack ranking or the normal distribution could suddenly become suitable. Their failings are built in, and supply chain leaders waste precious time and resources whenever they reach for these tools in the hope of a different result.
Negative knowledge is often far easier to grasp than its positive counterpart. One can spend days poring over how a columnar database functions, but it is much simpler—and more practical—to remember that such databases are unfit for systems of record. Any firm that runs a highly transactional workflow through a columnar architecture quickly discovers it runs painfully slowly and burns through computing resources, hiking hosting expenses. A company might endure this lesson only once before scrapping the entire setup, never repeating the misstep. In that sense, negative knowledge spares a manager from mastering every nuance of why a technology fails; it is enough to know it will fail, and to steer clear of that dead end altogether.
Conversely, negative knowledge is particularly suitable for supply chain because gathering trustworthy positive knowledge is so difficult.
Experiments in supply chain are deceptively difficult to conduct. No large company wants to gamble with the performance of a vast network whose failures can ripple through inventory, transportation, and customer satisfaction all at once. Even a seemingly modest trial can disrupt entire segments of the operation yet offer only a fleeting sense of whether it truly helped or merely shifted the burden onto another part of the chain. A retailer that experiments with smaller, more frequent deliveries might reduce local stockouts for a handful of stores, only to find that shipping costs surge and cause complications for distribution centers already juggling complex schedules. This is the recurring curse of supply chain: problems often get nudged around rather than eliminated.
Moreover, any experiment in supply chain, like an experiment in economics, is notoriously hard to replicate. No two moments in time offer identical conditions, and market pressures, organizational constraints, and the human factor can never be reproduced in full. What worked last year under a certain demand pattern might fail spectacularly when consumer tastes shift, or a new competitor enters the market. Confounding factors always lurk in the background, obscuring cause and effect. When a company declares an experiment “successful,” it is usually interpreting local data through a particular lens—often unaware that broader realities have already changed. Instead of guiding us toward conclusive truths, such trials mostly remind us how fragile and slippery those truths can be.
So-called “best practices” cannot be taken at face value. Professors, consultants, and software vendors each push their own agenda, and none are aligned with what a company genuinely needs. Professors focus on theories that lend themselves to research papers and classroom grading, neither of which has much bearing on whether those theories deliver real-world results. Consultants, for their part, clamor for originality so they can stand out among their peers, yet novelty alone proves nothing. By spotlighting fringe issues and framing them as urgent, they divert top management from far more pressing concerns. Meanwhile, software vendors relentlessly expand their feature sets to tick every box in an RFP, yet seldom acknowledge how these added layers push maintainability to the brink.
Even when a bright idea does find practical footing, it can lose its edge the moment too many adopt it. A so-called best practice, once widely diffused across an industry, no longer confers an advantage. Rivals adapt, eroding whatever gains the practice once yielded. This effect is not peculiar to supply chain, but it looms especially large here because so many competing players—vendors, consultants, even academic circles—are motivated to brand a new technique as a universal solution. Real progress lies elsewhere, for there is no single method so robust that it survives unchanged under the glare of mass adoption. What counts as a best practice depends on rivals not doing it, and the moment they catch up, its value vanishes.
The map of epistemic pitfalls is now complete: fashionable mathematics, story-driven sociology, corrupted evidence pipelines, and the perverse incentives that keep all three alive. Escaping this maze requires a reference frame that does not shift when fashions do, and a guardrail that keeps us from steering back into the same ditches. Economics supplies the frame; negative knowledge supplies the guardrail. Only by translating every supply‑chain claim into opportunity cost, profit, and loss—while actively rejecting patterns that predictably destroy value—can we compare alternatives on neutral ground and expose wishful thinking for what it is. With these standards in place, the next chapter leaves epistemology for first principles, revisiting the elementary laws of economics and showing how they anchor every subsequent decision rule in this book.
Economics
Supply chain is an applied branch of economics, itself a branch of praxeology, the theory of human action. Supply chains emerge from people acting in concert at scale to improve their material conditions.
In Chapter 1, we defined supply chain as the mastery of optionality under variability. Optionality reframes the “alternative uses” of general economics through a corporate lens: every decision allocates a scarce resource to one option while excluding the rest. Economics is not a distant backdrop—it is the grammar of supply-chain practice.
Economics unfortunately remains a science that is both largely disregarded and ignored. Across legacy and social media, many self-described “economists” speak not as scientists but as political ideologues16.
Conflicts of interest worsen matters: many economists draw their salaries, directly or indirectly, from the very states whose policies they comment on. Whether consciously or not, such funding narrows the range of views that can be voiced without professional risk17. When ideology masquerades as neutral analysis, the public’s distrust only deepens.
Essential economic principles
For readers unfamiliar with economics, we recommend Basic Economics (2000) by Thomas Sowell. While not a prerequisite, its insights are foundational to the study and practice of supply chain. Like Sowell’s other work, it is exceptionally clear and accessible to a non-technical audience. Above all, this book clarifies that economics is strictly neutral with regard to politics and other ideologies. The general laws of economics cannot be evaded: they apply equally no matter how society organizes itself; whether it be capitalism, socialism or feudalism. Laws such as supply and demand, diminishing returns, and Ricardian comparative advantage are fundamental to supply chain.
These fundamental laws govern every exchange that a supply chain orchestrates.
Law of supply and demand. In a free market, the price of a good rises when demand outstrips supply and falls when supply exceeds demand. The price signal is not an arbitrary convention; it is the device that rations scarcity. When inbound ocean containers were in short supply during the 2021 post-lockdown rebound, freight rates quadrupled. Shippers that valued the containers the most accepted the premium and shipped on time; lower-value loads waited until capacity returned. The mechanism can seem brutal, yet it silently assigns a scarce resource to its most valuable use while encouraging carriers to invest in extra capacity.
Law of diminishing returns. If one factor of production is increased while all others are held constant, the additional output generated by each extra unit eventually declines. Double the number of pickers in a fulfillment center and parcel throughput nearly doubles—until congestion at the packing stations begins to bind. The same curve appears when adding buffer stock: the first pallet slashes the risk of a stockout, the tenth pallet offers only a sliver of extra protection. Locating the knee of this curve is a central engineering problem of supply chain.
Ricardian law of comparative advantage. Even when one party is absolutely more efficient at producing every good, both parties gain from trade when each specializes in what it does relatively best. A Swiss micromachining workshop that makes high-precision gears and an Indonesian plant that weaves cotton shirts both prosper by exchanging output, despite wage and skill gaps. Supply chains furnish the infrastructure—shipping lanes, warehouses, information systems, and contracts—that turns comparative advantage into tangible surplus.
These three laws are illustrative, not exhaustive. Plenty of further concepts—opportunity cost, time preference, entrepreneurial profit and loss—enter the picture, yet this trio is enough to show why every stocking, routing, or pricing choice ultimately reduces to the same question: how best to direct scarce resources toward their highest-valued use under uncertainty.
Markets and supply chains
The remarkable improvement in humanity’s material conditions over the last three centuries can largely be attributed to the free-market economy, which has financed the development of nearly all the sciences and technologies we now enjoy. In developed countries—now covering an ever larger share of humanity—real prices of food and manufactured goods are a fraction of their levels a century ago.
The mastery of supply chain is one of the many innovations driven by free markets to deliver more with fewer resources. While supply chains rely on complementary technologies, their presence typically signals mass accessibility; without them, goods can still be produced, but at costs affordable only to a privileged few.
The superior material conditions in developed countries largely result from efficient supply chain operations. While advances in science and technology allow us to achieve extraordinary feats—such as sustaining life in orbit—there is a vast difference between isolated technological stunts and the packaged, reliable miracles accessible to the masses. This gap is largely due to robust supply chains, which serve as the backbone of our industrial civilization.
A revealing counterexample comes from the Soviet Union. Under central planning, almost every large manufacturer was forced into extreme vertical integration because reliable partners simply did not exist. A refrigerator plant in Leningrad ran its own timber yard for crates, its own glass foundry for shelves, and even a dairy farm to feed the company canteen. What looked on paper like enviable self-sufficiency was, in fact, the symptom of an economy where trust, specialization, and price signals had been extinguished.
Lacking markets that rewarded punctuality and quality, “partners” were merely other ministries obliged to deliver whatever they could, whenever they could. Factories therefore hedged by duplicating upstream capacity in-house and by hoarding months—or sometimes years—of input stock. Finished goods fared no better: they often waited weeks for railcars, themselves hostage to the chronic wagon shortages of Gosplan’s schedules. Lead times that would have bankrupted a Western firm were accepted as routine, and the waste showed in every metric. Ton-kilometers of freight per unit of output were several times higher than in market economies, while specific energy consumption for steel, textiles, or bread dwarfed Western benchmarks.
This episode matters for present-day managers: supply-chain performance is not chiefly a matter of clever heuristics or larger warehouses. It is first and foremost a matter of incentives coupled with adequate economic calculation. When prices move freely and contracts are enforced, specialization flourishes and long, delicate networks become not only feasible but vastly more efficient than monolithic giants. Remove those incentives or that calculation, and the networks collapse inward until every plant becomes a fortress hoarding its own raw materials. Complexity does not disappear; it merely resurfaces as waste.
Before the 1950s, air travel was an extravagant luxury reserved for the military and the wealthy. A series of innovations then made flying safe and affordable, rendering it commonplace. A key innovation was podded engines—aircraft engines mounted under the wings in pods—a concept pioneered by the Sud Aviation Caravelle in 1955. These pods decoupled engine maintenance from that of the main fuselage, substantially lowering operating costs and paving the way for aircraft like the Boeing 707 in 1957 to democratize air travel.
Currently, the space industry has yet to benefit from modern supply chain practices, mainly due to technological limitations. A handful of wealthy entrepreneurs have bought flights to orbit, though the prices are still prohibitive for most. After decades of stagnation, SpaceX is advancing technologies—most notably reusable rockets—that could eventually support a space supply chain. Once these technologies are mastered, established supply chain principles will lower costs substantially.
A historical example of this dynamic is the rapid adoption of containerization in maritime shipping during the 1950s and 1960s. The standardized intermodal container offered a breakthrough that drastically reduced handling costs and transit times. Once this technology was established—requiring suitably equipped ports, cranes, and ships—a new generation of supply chains formed around it. Ports that failed to upgrade and integrate container standards lagged behind, while those that embraced containerization became major global trade hubs. Likewise, whenever a new technology matures—whether it be reusable rockets or advanced robotics—innovators quickly organize their supply chains to exploit these advances, leveraging both modularity and standardization to achieve unprecedented scale and cost efficiency.
More generally, technological advances expand the range of available options, and supply chain decides how best to leverage them for maximum economic efficiency.
The value of supply chains
Supply chains emerged from the market economy because they are efficient. Dating the first true supply chain is vain: the phenomenon emerged gradually and can be traced back to antiquity. Moreover, supply chains emerged without visionary planners deliberately steering companies toward what we now recognize as supply chains. Companies simply followed empirical efficiency, and it led them to the supply chains that exist today. While modern companies can now be engineered with the explicit intent of bringing a superior supply chain to market, this perspective is fairly recent—and largely posterior to the supply chains themselves.
Supply chains exist because they unlock a handful of structural mechanisms that confer durable competitive advantages. The pages ahead examine these mechanisms and show how supply-chain design can generate more value than sheer size ever could.
These mechanisms differ somewhat from the traditional concept of economies of scale, generally understood as achieving greater efficiency with increasing size. That efficiency typically arises from the mix of fixed and variable costs in production. Functions such as engineering, marketing, strategic management, and compliance are largely fixed regardless of output, whereas acquiring, transporting, and processing goods scale with volume.
Economies of scale represent a basic form of flow optimization: increasing volume reduces marginal cost. In everyday operations, however, supply chain professionals rarely have levers to raise volume enough for additional economies of scale; step-changes in flow belong to M&A (mergers and acquisitions), not supply chain.
Beyond economies of scale, additional value stems from how a company shapes its flow through four recurring forces: standardization (taming innate variability), pooling (consolidating inventory to buffer randomness), batching (grouping movements to cut unit-handling costs), and versatility (designing assets to serve several purposes). These forces pull on different levers—information, space, time, and capital—yet converge on the same goal: delivering more service from the same scarce resources. The next sections examine each force in turn, starting with standardization.
Standardization
Variability is the natural state of physical things. No two trees are exactly alike; the same holds for stones, cows, people, and parcels of land. Boundaries often blur: do fallen fruits still count as part of the tree? Does the ownership of a piece of land extend from the depths of the earth to the stars above? Our actions, too, exhibit inherent variability. Even arranging an apple admits countless variants; no two placements are exactly the same. The apple may fall on the floor, be picked up again, and put back on the shelf. Does this still count as the same action? Even if the floor is dirty? Yet in industrial civilization we take for granted that, within the flow of physical goods, almost every “thing” and “action” can be neatly identified and counted. This is no accident; it is the result of centuries spent making the flow manageable—a process called standardization.
Standardization is a loose collection of practices that make the flow of goods manageable within firms and across markets. Indeed, inherent variability in all things complicates everything. Every item purchased or sold varies in its characteristics and, consequently, often in price. Every manufacturing process must be continuously adjusted to accommodate the minute differences in its ever-changing inputs. These complications consume managerial attention—a costly input—and blunt the company’s capacity to achieve economies of scale. They also impose cognitive costs on customers, who must judge whether an item suits its intended use.
To a present-day reader, this may seem abstract. The vast majority of goods are standardized—mass-produced within narrow physical tolerances. Variations between units are expected to be imperceptible; two units of the same product should be visually indistinguishable. When such a feat is possible, we, the customers, instinctively attribute the difference to a defect of some kind18. Even store-bought fruits and vegetables are now carefully calibrated, reducing variability to a fraction of their natural differences at harvest. In industry, quality chiefly means the absence of unintended variation.
We also expect high modularity and interoperability among assets. Containers flow seamlessly from ships to trucks, thanks to dedicated container cranes. A limited set of tire sizes covers nearly all trucks. Standard pallet patterns optimize container space, and forklifts—built for pallets—move them efficiently. At every stage, barcodes are printed and later scanned—often by another company. Ultimately, modularity and interoperability stem from careful standardization, which simplifies entire classes of operations.
Yet, just a century ago, the situation was radically different. Most mass-produced items still exhibited perceptible variation. Interoperability was local, not global, and modularity largely confined to individual companies. After World War II, standards expanded in scope and depth alongside modern supply chains. The movement had forceful advocates, notably W. Edwards Deming, whose quality methods made standardization a managerial priority.
Standardization is a critical requirement for massifying the flow: without it, many processes are infeasible, brands are exposed to defects, managers chase outliers, and most economies of scale remain inaccessible.
As a corollary, the flow has become predominantly discrete: itemized units—counted one by one—now travel through the network instead of anonymous tons or liters. Bulk modes persist for sand, crude, and grain, but for anything with a hint of value, the default is a barcoded, scan-ready unit.
This discreteness supercharges optionality. A railcar of ethanol presents one choice—ship or hold—whereas a pallet of 120 power drills presents 120 separate choices: every drill can be split, rerouted, repriced, kitted, or postponed. Each SKU thus functions as an option ticket whose value depends on where, when, and to whom the option is ultimately exercised.
The upside is obvious: fine granularity lets a company redirect stock to emerging pockets of demand, arbitrage tiny price spreads, and delay commitment until the latest possible moment. The downside is equally clear: the decision space explodes. A planner who once scheduled two containers a month must now orchestrate thousands of micro-moves a day. When cognitive or computational bandwidth is lacking, management falls back on coarse rules—MOQs, fixed cycles, blanket allocations—thereby erasing much of the flexibility it paid for. Discreteness is thus a double-edged sword: it multiplies opportunities while magnifying the need for automation capable of scanning and ranking them. We will return to this tension when discussing decision-making processes.
Handling individual units is not cost-free—the carton, the barcode, the extra touches all add pennies—but for anything above the cheapest commodities, those pennies are dwarfed by the upside of precise allocation. In the pages that follow, every unit will be treated as a negotiable option, and the question will always be which exercise maximizes long-run profit. With this discrete foundation in place, we can now turn to the second structural force: pooling.
Pooling
Pooling concentrates inventory in one place—or under one corporate entity—to serve multiple demand streams. If the overhead from concentration is negligible, pooling delivers the same service with less capital than multiple independent inventories. In practice, pooling is profitable when the inventory cost savings exceed the additional service overhead. For example, concentrating inventory in one place typically raises transportation costs relative to serving goods from multiple locations.
Pooling mitigates supply-chain variability by the law of large numbers: a central pool can flexibly absorb uneven surges across segments, reducing the total stock needed to maintain service levels. Imagine one table holding champagne for the entire reception rather than many smaller tables. Pooling resources reduces the risk of any one table running dry, ensuring that everyone is served. Of course, pooling can hinder last-mile service if it places goods too far from consumption.
Consider a company weighing a single large distribution center against multiple smaller warehouses. A single center pools demand efficiently and reduces total stock, but it also adds overhead—longer average hauls and more handling at the hub. Managers must weigh these logistics costs against working-capital savings. If regional product demands are closely correlated—due to seasonality or nationwide campaigns—a single warehouse delivers less benefit. Correlated demand and customer flexibility to switch stores or channels erode theoretical pooling gains. Side-by-side simulation or a pilot usually tests whether inventory savings outweigh added expenses and delays, ensuring a positive net value.
In theory, pooling can dramatically reduce working capital; in practice, frictions pare back the gains and add overheads. Even so, pooling remains attractive in many contexts.
Putting transportation overheads aside, the gains are lower for two reasons. First, the streams of demand are never independent; on the contrary, those streams should be expected to be quite correlated. Markets are connected; while offsetting surges and drops can occur, the typical pattern is co-movement, not cancellation. Correlation has many sources, often the company’s own actions. For example, a nationwide advertisement raises demand across many locations simultaneously. If streams are perfectly correlated, pooling requires as much working capital as decentralization for a given service level.
Second, when demand surges, a customer facing a local stockout may seek service from another location, incurring the associated overheads. This is common in dense retail networks where multiple stores compete within the same city. Moreover, letting customers check availability online amplifies the effect, nudging them—up to a point—to cooperate with the company’s stocking policies. If customers fully accept the overhead of alternative service locations, pooling brings no benefit: the same inventory reduction was achievable without consolidation.
If the future were perfectly anticipated, pooling would confer no benefit; it would only add overhead from suboptimal placement. Pooling is widespread precisely because future uncertainty is irreducible.
In terms of optionality, pooling reduces the number of options but neither their frequency nor their diversity. Thus, pooling somewhat simplifies the supply chain challenge. However, firms rarely adopt pooling for the simplification alone. Empirically, concerns about dead stock and shelf space dominate the selection of locations suitable for storing a given product.
Batching
If pooling concentrates inventory in space, batching concentrates the flow in time. Batching processes items in larger lots rather than singly, lowering per-unit processing cost but reducing the ability to serve small quantities.
Batching is ubiquitous in logistics: once the flow exceeds a threshold, units are repackaged into larger lots. A typical progression is: group $$N$$ units into a box; bundle $$N$$ boxes into a pallet layer; combine $$N$$ layers into a full pallet; finally pack $$N$$ pallets into a container. Bundling straightforwardly reduces handling costs, both within a facility (intralogistics) and between facilities. Bundles simplify handling and counting, and their packaging often adds physical protection.
Batching is also common in manufacturing: it typically means running a large lot of a single product before switching or pausing. This approach minimizes setup time and other non-productive work (e.g., recalibration) and also facilitates post-production logistics. In process industries (e.g., chemicals), scale effects such as higher yields push toward enormous plants and correspondingly large batches.
Batching increases the inherent variability of the flow and amplifies the effects of external market fluctuations. Internally, variability rises because batching ties the fate of many units together. If 100 units are shipped individually by various couriers, it would be surprising for all to be lost. However, if all 100 units travel on one truck, the (small) traffic accident risk now applies to all at once. Formally, batching raises variance in the flow.
Batching amplifies external variability. The most frequent mechanism is a longer flow cycle. Batching reduces the frequency of acquisition, transport, and production. Colloquially, this extended cycle is called “longer lead time”. In batching, lead time is the period the company must wait before launching the next batch. Larger or less frequent batches tie up more working capital in inventory because items must be purchased or produced well in advance. They also diminish flexibility: misjudged demand turns oversized batches into overstocks or prolonged stockouts, both costly to correct. A longer cycle forces the company to operate longer before it can decide again. For example, with batching, misjudged demand enlarges overstocks or lengthens stockouts, depending on whether demand was overestimated or underestimated.
Batching lowers optionality by reducing both the frequency and the number of decisions within a period (e.g., a month). In the absence of adequate automation, batching is often employed primarily to simplify the supply chain. But this trade-off leaves the company with fewer options and less flexibility to adapt to market variation. In practice, companies too often opt for large batches because they lack the capacity or managerial resources to handle more frequent, detailed decisions. This reflects a human or organizational limitation, not a physical necessity.
With proper automation, the overhead of frequent decisions about physical goods becomes minimal, often negligible. As automation lowers the cost of orchestration, the risk–reward of smaller batches improves, reducing both inventory costs and lead times. Indeed, while modern computers operate on a nanosecond scale, physical goods are limited to timescales of tenths of a second (0.1 s) – a disparity that accommodates the slower pace of physical processes without causing damage. Consequently, to a modern computer, the flow of physical goods appears almost frozen19.
Versatility
Versatility denotes capital goods—durable assets such as machines, vehicles, or facilities that enable the flow of physical goods—that can serve multiple distinct functions or even an entire spectrum of operations. Yet at any moment a capital good can serve only one purpose; thus the company must assign each capital good a specific purpose at that time.
For example, a warehouse designed to maintain a controlled room temperature (e.g., 15°C–25°C) might also be equipped with devices capable of supporting a cool chain (e.g., 12°C–14°C). Even if the cold-chain capability remains unused at first, it can be a sound investment to accommodate future market fluctuations.
Versatility is deliberate: it enhances the company’s agility and expands its range of options. It always entails overhead associated with its wider capability set. All else equal, a versatile capital good costs more than a specialized one because of its broader capabilities. In practice, these overheads are usually offset by modularization—and its frequent companion, standardization.
Modularization breaks a process into smaller, interchangeable units that can be easily reconfigured for different tasks. Once standardized, companies aggressively compete to produce these units. This typically yields significant economies of scale, leaving companies little choice but to adopt modular capital goods to stay competitive. A downside is increased overall flow complexity as operations are subdivided merely to fit widely available modules.
Modern logistics exemplifies modularization in its barcodes, loading docks, and shipping containers. Without the intense standardization of the 20^th^ century, the benefits of modularization would be far smaller. A company can print a barcode and expect another to read it because they share a standard. A truck can be loaded at one site and unloaded at another because dock heights match the truck’s design.
Modern manufacturing has likewise transformed. Companies now leverage standardized units for most processes, resorting to custom-made capital goods only for steps that offer a unique competitive edge. Often, the entire manufacturing setup consists of standardized components, so a company’s distinctiveness stems from its unique selection and arrangement of these units. Moreover, the steady progress of programmable controllers—in various forms—has rendered these components not only modular but also highly versatile.
Every versatile capital good requires its own allocation schedule. At any given moment, the company must decide whether to reassign an asset to a different purpose. For some assets, this decision is straightforward and embedded in practice—for example, truck fleets typically operate on fixed routes. Conversely, some assets’ versatility remains untapped; a facility might produce a narrow range for years even as market conditions fluctuate.
The versatility of capital goods vastly inflates the number of options available to the company. Not only can alternative goods be served with the same apparatus, but the same service (transport or transformation) can be performed many ways, as versatile capital goods often overlap in capability. For instance, box trucks and semi-trailers are nearly interchangeable in transporting goods; only the heaviest loads require a semi-trailer.
Yet each asset carries distinct acquisition and operating costs. Furthermore, every asset, whether labeled perishable or not, will in time perish through wear and tear. Consequently, much is at stake in deploying the right asset at the right time for each task to achieve the outcome while minimizing costs.
From forces to policies
Viewed through the economist’s lens, these forces—standardization, pooling, batching, and versatility—are four ways to reshape the frontier where scarce resources meet alternative uses.
Each force corresponds to an elementary economic trade-off. Standardization lowers marginal cost at the price of variety; pooling frees working capital by accepting longer first-mile hauls; batching swaps inventory risk for lower handling cost; and versatility converts capex into real options whose value materializes only if the decision system can pick the right moment to exercise them. Together they expand the option space far faster than managerial attention can track.
Faced with a combinatorial explosion of options, companies often adopt policies expressed as a simple allocation rule. For instance, a company with two warehouses and a fleet of trucks might decide that one truck should make a daily round trip between the facilities to rebalance inventory. This policy is adopted with the implicit intent to use the rigid schedule for inventory rebalancing. The goods to be transported are unspecified, but the tacit understanding is that managers will use the daily allotted capacity to keep inventory balanced. However, this “one truck” policy is simplistic; some days, two trucks would be preferable, while on others, minimal inventory imbalances would justify postponing the round trip. Such rule-based policies may be implemented through software to notify the right employees at the right time, but a clerk could have otherwise supervised them with little effort.
However, governing the supply chain through simple rule-based policies is largely vestigial. It reflects constraints that ceased to be relevant with the advent of digital supply chains. Sticking to simple rule-based policies was essential when a clerk was in charge of executing the policy. However, if the policy is delegated to a computer, then it can be made much more dynamic while keeping the processing overhead negligible. If this added complexity improves the supply chain, a far superior allocation of resources follows. Revisiting the example, a more advanced policy should not only consider the amount of inventory to rebalance, but also compare the benefits of rebalancing with the alternative uses of the trucks, such as serving customer orders. If trucks are the bottleneck and the inventory imbalances between the two warehouses are tolerable, all trucks should be allotted to customer deliveries.
Unlike pooling and batching, which reduce the number of options, the versatility of capital goods inflates them. There is no a priori reason to think that all those options are economically equivalent. On the contrary, many options are poor, and one is best under a given economic model of the supply chain. Software supported by modern computing hardware makes it relatively straightforward to search through an extremely large number of options. The fine print of this process is unimportant here; what matters is that versatility is profoundly consequential for supply chains.
Because every policy is ultimately an allocation rule, its merit is economic, not procedural: it must raise the risk-adjusted rate of return of the resources it steers. Static heuristics were defensible when humans were the computational bottleneck, but once a computer owns the rule, the benchmark becomes opportunity cost, not clerical convenience. Automation therefore matters only insofar as it lets the price signals—expressed inside the model as valuations and discount rates—propagate to every micro-decision. In the end, a policy is justified exactly to the extent that it generates profit.
In practice, the combinatorial explosion caused by versatile capital goods can appear daunting and often easily outstrips human oversight. Yet modern supply-chain software does not simply attempt to brute-force every possibility. As will be discussed in Chapter Decisions, advanced analytics—supported by solvers and heuristic algorithms—can render this complexity tractable.
The goal of supply chain
Supply chain was defined in the first chapter as the mastery of optionality in the presence of variability when managing the flow of physical goods. Mastery here is not an abstract virtue: it is the concrete ability to turn optionality into money while shielding the company from the downside of uncertainty. In a market economy, the yardstick for such mastery is unambiguous—long-run profit expressed in hard currency. The practical goal of supply chain practice is therefore to raise the firm’s risk-adjusted rate of return on every scarce resource it touches—capital, capacity, time, goodwill—by continually improving both the range and the quality of the options that can be exercised.
This goal subsumes every secondary intention a practitioner might entertain. Higher service levels, shorter lead times, leaner inventories, greener transportation, happier suppliers, safer working conditions—all matter, but only insofar as they contribute, directly or through risk mitigation, to the net present value of the business. Supply chain is not moral philosophy; it is an applied branch of economics that survives or perishes by the ledger. Pursuing profit must not be confused with myopic cost cutting or the caricature of a “naked” profit motive; it is the disciplined quest for the allocation that maximizes long-run surplus while respecting both variability and optionality.
The remainder of this section unpacks that proposition. First, we revisit profit and loss as the indispensable feedback loop that tells a company whether its supply-chain decisions create or destroy value. Next, we explain why objective valuations are mandatory inside the firm even though all valuations are subjective at the individual level, and how the notion of rate of return makes disparate decisions comparable across time. Finally, we address frequent objections—moralistic, apocalyptic, or institutional—to profit seeking and show why none of them exempts a supply chain from the laws of economics.
Profit and loss
Profit and loss is not an accounting gadget; it is the scoreboard of a society built on consent rather than coercion. Whenever two parties trade, each expects to be better off—otherwise no deal occurs. Profits, therefore, arise only within the domain of voluntary exchange. Conversely, whenever allocation is imposed by force—or by regulation that mimics it—the very possibility of profit vanishes—replaced by tribute, tax, or plunder20. In economic terms, human relations lie on a continuum from hegemonic command to unfettered trade. Profits arise only on the latter side.
Yet a vocal strand of public opinion still portrays profit as intrinsically exploitative21. This view conflates profit with plunder: the latter is taken by force, whereas the former can arise only when both parties consent and deem themselves better off after the exchange. Seen this way, profit has a moral edge: it is earned solely by serving others. Losses play the mirror role. They warn entrepreneurs that customers prefer other uses for their scarce resources. Far from being damaging, losses act as the market’s immune system; they halt further waste and redeploy labor, capital, and ingenuity to more urgent purposes.
History makes the contrast vivid: before the commercial revolution of the 17^th^ century, fortunes were typically amassed by conquering a province, taxing the peasantry, or courting royal favors. Since the rise of open markets, the road to wealth has inverted: Rockefeller refined kerosene more cheaply than anyone, Akio Morita miniaturized electronics for the masses, Ingvar Kamprad taught flat-pack furniture to a frugal planet. Where profits are contestable, efficiency soars; where allocation is decided by decree, queues, shortages, and rust follow.
Supply chains translate this principle into pallets, containers, and transport lanes. Standardization shrinks unit cost, pooling frees working capital, batching arbitrates handling expenses against responsiveness, and versatility turns fixed assets into spring-loaded options. Yet none of these levers tells you when to stop. The only impartial referee is the profit and loss statement: does the extra SKU, the larger batch, or the reinforced crate pay for itself once variability and opportunity costs are tallied? If the answer is yes, the flow expands; if the answer is no, the flow contracts and capital migrates to better uses.
Practitioners, therefore, miss the point when they chase “service level”, “inventory turn”, or any other local metric in isolation. Such indicators matter only insofar as they raise the firm’s long-run rate of return. The objective never changes: maximize the gap between what customers willingly pay and what supply chain irrevocably consumes, subject to the risks imposed by uncertainty.
The sections that follow dismantle the stock objections to the profit motive and show how a clear P&L perspective inoculates supply-chain design against both spreadsheet sophistry and pious slogans.
Objective valuations
Economics begins with a simple observation: only individuals can feel, think, and act; therefore, only individuals can ascribe value to anything. These valuations are subjective and ordinal—one can say he prefers coffee to tea, but there is no arithmetic to add, subtract, or average the two preferences. Worse for the analyst, the ranking can flip tomorrow without warning. No spreadsheet can pin down an “inherent” price for the first sip of water, the hundredth, or the thousandth; the paradox of value remains a feature, not a bug, of human action.
Statistical models can, of course, learn to predict how much a person of average means will pay for an espresso or how likely he is to abandon a shopping cart, yet the forecast does not turn a private preference into an objective quantity. It merely characterizes the subjectivity as a probability distribution, useful in commerce but silent on worth.
This point torpedoes the labor theory of value, which tried to anchor prices to the number of work hours “embodied” in a good. Diamonds are dear not because miners toil, but because buyers prize the sparkle. If tomorrow everyone decided that quartz is prettier, diamond mines would shutter despite unchanged labor inputs.
In 2019, when a massive fire broke out in Notre-Dame, the public was swiftly evacuated. While no human lives were at stake anymore, some firefighters put their lives at great risk to save what they considered timeless works of art. In economic terms, they exercised a value judgment, placing mere artifacts above their own lives. Most men, in the same situation, would not have acted likewise and would have stopped risking their lives once it was clear that no further lives were endangered. It is pointless to assess how much those artifacts are “worth” in hours of labor, and how much it would cost to produce near-perfect replicas. From the perspective of those men, the historical and artistic value of those artifacts was categorically superior to any later attempts at replacing them. Through their actions, free individuals can exercise entirely arbitrary value judgments. The role of economics is not to assess whether such value judgments are good or bad; it merely recognizes their irreducible subjectivity.
Companies, however, cannot indulge such idiosyncrasies. Inside a firm, the CEO may love truffles and the CFO may hate them, yet the only admissible question is: “Do customers pay enough, soon enough, to cover our costs?” Individual preferences are not additive; they cancel out. What remains is the impersonal verdict of the market, recorded as money in, money out.
If an employee decides that his company must value certain options disregarding the profit motive—“we should stock this eco-friendly widget even if it never sells”—this employee is not being good or caring; he is just deluding himself by projecting his own value judgment onto the company. By doing this, he is doing a disfavor to the company, and to the market, because he is acting against the wants of the consumers, favoring instead his own pet perspective of the day. Naturally, if the employee disagrees with the general mission of the company, then it is his duty, as a free individual, to resign and seek employment elsewhere. Disagreements with the company’s stated purpose are not to be settled by sabotaging operations from the inside.
A firm is therefore an instrument of social cooperation: it aggregates capital, talent, and information to satisfy the most urgent wants that other people express through their purchases. Profits and losses turn those dispersed signals into an objective yardstick that every planner can read: more coins earned than spent means resources flow toward higher-ranked wants; the opposite means they are wasted. Supply-chain practice must anchor every valuation to that yardstick—never to personal taste—because only prices let millions of subjective scales collapse into one objective measure.
The next question, then, is how to compare two profitable options that differ in timing—as with rates of return.
Rate of return
Whenever a decision separates committing a resource from its payoff, plain profit arithmetic no longer suffices. Economists call this universal phenomenon time preference: identical goods are worth more the sooner they are available. A cup of coffee served now is worth more than the same cup promised next week; a supplier usually demands a premium for payment deferred by ninety days.
Time preference is first an individual trait. Each person privately discounts future benefits according to circumstance, and the market aggregates those discounts into the interest rate. The interest rate therefore quantifies, at the scale of society, how much present goods are preferred to future goods.
Corporations inherit the same logic. Cash received today can be reinvested, used to settle liabilities, or serve as a buffer against uncertainty; cash received next year cannot. Firms therefore embed an explicit cost of capital—their collective time preference—into every investment test. However appealing in isolation, any proposal that fails to clear that hurdle erodes value once the opportunity cost of waiting is recognized.
Once time preference is acknowledged, optimizing profit requires attention to time; specifically, the company should pick the options with the highest rate of return. Conceptually—setting aside secondary complications—the rate is computed as
$$ \mathtt{RoR}=\frac{\text{revenues} - \text{spending}}{\text{spending} \times \text{period}} $$
where period is the time required to realize the associated revenues and spending. The rate of return (RoR) is expressed as a percentage per unit of time. If the RoR is greater than zero then the company operates profitably; if the RoR is less than zero then the company operates at a loss. When comparing two options, prefer the higher rate—provided both are normalized to the same period.
For example, let’s consider a retail store that can choose between two products to be put on display on a given shelf. Both products are sold at 3 coins with an average demand of 1 unit per day. The first product costs 1 coin per unit, while the second product costs 2 coins per unit. Storage costs, setting shelf space aside, are considered negligible. Also, the shelf is large enough to hold enough units to make stockouts negligible. In this situation, the first product has a RoR of 200% per day, against 50% per day for the second product. Thus, the first product should be preferred because it has the higher RoR. This result is unsurprising: the products are identical except that the first has twice the gross margin of the second. The RoR simply provides the ranking criterion that formalizes this choice.
In typical supply chain settings, maximizing RoR tends to allocate more capacity, stock, and shelf space to products with higher volumes or faster rotations. However, RoR—not arbitrary noneconomic metrics such as service levels—must be optimized. Many supply chain authors mistake the loose correlation between higher RoR and higher volumes or rotations for causation. As a result, inventory classification methods are frequently proposed to segment products into classes22 reflecting their volumes and/or their erraticity and adjust the resource allocations according to this segmentation. These methods confuse correlation with causation: RoR must be optimized. Mimicking its loose consequences is a categorical error that yields diminished returns.
In practice, the rate of return presents a series of complications:
-
Calculating RoR depends on revenue and spending attribution—a difficult problem.
-
Revenues and spending do not occur at the same time; a shared period rarely exists.
-
Revenues should occur as early as possible and spending as late as possible; cash flows must be discounted.
-
The lag between future spending and future revenues must respect the firm’s liquidity constraint.
-
Dependencies between options make RoR path-dependent on the picking order.
-
Both future revenues and spending are uncertain, and this uncertainty must be modeled.
None of these complications prove insurmountable in practice, even if they all entail some degree of approximation.
Revenue and spending attribution
RoR reflects an idealized view of physical flows. However, constraints require nontrivial economic modeling. As a result, in order to compute the RoR of a given option, the option must be disentangled from the firm, typically through economic attribution. This point is discussed below; for now note there is no one-to-one mapping between revenues, spending, and options. As a result, RoR is a modeling perspective on supply chain, not a self-evident truth.
For example, consider a retail store with two alternative products. The first product requires 100 coins of inventory—renewed daily—to generate 140 coins of daily revenue. The second product requires 150 coins of inventory to generate 200 coins of daily revenue. The two products take exactly the same shelf space in the store. In this case, the RoR of the first product is 40% per day while the RoR of the second is 33%. Thus, by the principle of picking the options yielding the highest RoR, the store owner should pick the first product. However, this choice is surprising: it selects the product that will ultimately generate less absolute gross profit over time. For a store dominated by fixed costs, this seems incorrect.
Indeed, the naive model above ignored the store’s fixed costs. Let’s refine the model by attributing a portion of fixed costs to the products. An attributed daily fixed cost is 40 coins. Thus, the daily cost for the first product becomes $$100+40=140$$ coins, and for the second $$150+40=190$$ coins. Under this revised model, the RoR of the first product becomes 0% per day while the second becomes 5.3%. Therefore, under this refined model, the second product should be favored.
This situation is not exceptional: what qualifies as best depends on the adopted economic model for the supply chain. Unfortunately, no single, straightforward criterion decides which model is superior; adequacy hinges on how well each captures the economic reality it is meant to inform. This challenge is a problem of general intelligence. That does not mean anything goes; we revisit this class of challenges later on.
Aperiodic rate of return
In finance textbooks the rate of return (RoR) is almost always tied to a fixed horizon—a month, a quarter, a year—because that is the cadence of accounting and taxation. At the micro level of supply-chain execution, however, such calendars are an awkward fit. Raising a purchase order, launching a markdown, allocating a pallet, or scheduling a truck are atomic decisions whose cash outflows and inflows trickle in at irregular intervals. A twelve-month yardstick blurs the very reason they exist: to recycle capital as fast as possible.
The aperiodic RoR removes the calendar straitjacket. Instead of forcing every option into an arbitrary window, it asks a single question: after how long will the option have repaid the cash it tied up and started compounding its surplus? That self-selected delay becomes the implicit period in the RoR formula, yielding a compound rate comparable across decisions, even though each chooses its own clock.
Because the computation is performed at the level of individual options, the spread of results is wide—differences of one and even two orders of magnitude are routine. A fast-moving, high-margin refill can clear 1,000% per annum once expressed in aperiodic terms, while the replenishment of a slow-seller in the next aisle may struggle to reach 10%. Such dispersion is not a modeling flaw; it is the signal that tells capital where it is urgently needed and where it quietly dies.
Operationally, the aperiodic RoR slides a cursor along the option’s cash-flow timeline, computing at each date the rate that would prevail if the position were closed out then. It keeps the maximum value; the date that yields this peak becomes the option’s implicit horizon. In one number, the metric captures both the scale of the gain and the fastest pace at which the tied-up capital can compound. The exact rule is spelled out in the annex.
Discounted cash flows
The previous subsection introduced the rate of return as a yardstick for how fast committed capital compounds. Speed, however, is only half the temporal story. A coin received today is worth more than the same coin received next quarter, because the earlier coin can be reinvested, held as a buffer, or simply earn interest in the meantime. Discounted cash flow is the numerical translation of the time preference: each future inflow or outflow is multiplied by a discount factor that shrinks with time, so every cash movement is expressed in “today-coins”. The discount factor derives from the firm’s cost of capital—the interest effectively paid for liquidity. For the precise formula, see the annex.
Put differently, discounted cash flow measures the absolute amount earned (or lost) while accounting for the time value of money. From the supply chain perspective, the interest rate represents the company’s financing cost. In finance, this rate often also proxies uncertainty associated with delays, but in supply chain use cases uncertainty is better handled through other mechanisms detailed later.
A practical illustration of discounted cash flows in shipping and stock decisions appears in fashion, which often faces long lead times and cash constraints. Imagine a mid-tier fashion brand that must place orders for its new collection six months in advance of the season. The brand negotiates terms requiring a substantial deposit upfront, with the remainder due upon shipment. Even if borrowed funds cost only, say, 10% per year, the brand’s effective rate of return on these early orders can dwarf that figure. This is because the brand’s capital is tied up long before revenues materialize, and any misjudgment in sizing, color, or style inflates the effective cost of the goods. When the season arrives, the brand may discount slow-moving items, adding to the opportunity cost of money that could have been invested in a more promising line. Conversely, a well-placed order that aligns with demand can rotate quickly at full price—delivering a rate of return that far surpasses the cost of capital. Thus, the interplay of long lead times, extended payment terms, and market uncertainty often outweighs the ambient interest rate in determining the real profitability of each supply chain decision.
In practice, for supply chain decisions, the ambient interest rate matters less than the rate of return, largely driven by inventory rotations. This is not to deny the importance of interest rates. Yet a unit of inventory yielding 10% per rotation and turning ten times per year compounds to $$1.1^{10}\approx 2.59$$, i.e. 159% per year. A single company often sells products whose rates of return range from 1% to 1,000% per year. This wide spectrum stems largely from inventory rotations that range from daily to annual. As a result of such extremes, the profit-maximizing quality of service varies greatly from one product to the next. Stocking strategies diverge accordingly.
An annual rate of return of 1,000%, while not atypical for the flow of an isolated product within supply chain, exceeds by more than an order of magnitude what finance deems “plausible”. Naturally, this calculation ignores fixed costs—above all, infrastructure. Once those costs are accounted for, effective rates of return in supply chains align with those in the general economy.
Cash flow and working capital
Interest rates set the floor for the cost of liquidity; they therefore matter to every supply chain. In practice, however, they usually hover within a relatively narrow corridor—say roughly 10–20% per annum—whereas inventory-cycle returns routinely span two orders of magnitude. Consequently, what most often constrains planners is not the precise price of money but its availability. A buyer who can draw cash at 12% instead of 8% will still proceed if the stock he finances compounds at 300% per year; conversely, a nominal 2% loan is useless once the credit line is exhausted. Cash-flow stewardship is therefore critical—especially in businesses that must wire funds months before the goods clear customs or reach the shelf; fashion collections, holiday toys, and high-tech gadgets are everyday examples.
Take, for instance, a firm whose bank extends a revolving credit facility capped at 25% of the company’s average turnover over the past three years, charged at a flat 10% per annum. Up to that limit, cash can be drawn overnight; the instant the cap is reached, however, every extra coin requires a fresh negotiation and is subject to refusal. Alternative lifelines—equity injections, mezzanine debt, supplier factoring—certainly exist, but none can be secured overnight, and all carry significantly higher explicit and implicit costs. Liquidity is therefore priced by a piecewise curve rather than a single number, and any model that assumes one uniform interest rate ignores precisely the nonlinearity that most often bites in practice.
The nonlinearities associated with access to liquidity explain why payment terms are often of prime importance in supply chain. Those payment terms critically shape the amount of working capital required by the company. In the long run, lower working-capital requirements improve profitability by reducing interest paid to lenders; in the short run, they often make the difference between a company that operates smoothly and one forced into inferior choices because better options are foreclosed by insufficient liquidity.
For example, a fashion brand may have to place purchase orders for pieces in its new collection more than six months before the season starts. Overseas suppliers may require substantial up-front payments. If the fashion brand anticipates stronger sales for this collection, it may issue larger purchase orders. However, the company may be blocked by an insufficient credit line. In this case, smaller purchase orders follow, possibly causing substantial stockouts downstream. Naturally, from the lender’s perspective, the cap is not arbitrary: the lender simply does not share the brand’s optimism about the collection’s outsized success.
The flat-interest-rate assumption rarely holds, though there are situations where it is an acceptable approximation. The simplest case is when the company has access to liquidity that dwarfs profitable supply chain opportunities. A slightly more subtle case is when the company has negative working-capital requirements.
To illustrate, consider a retailer that keeps store inventory turning monthly with negligible supplier lead time. Assume suppliers are paid two months after delivery, and products are sold in the store with a 20% gross margin. Under those conditions, for every four coins’ worth of goods held in store, the retailer has $$-2$$ coins of working-capital requirement. Indeed, by the time the goods must be paid, the inventory will have rotated twice on average, yielding two coins of gross profit. Here the retailer has negative working-capital requirements—at least from the perspective of cash flow alone.
This example must not be confused with negative interest rates; the retailer remains subject to market rates.
Returning to the more typical case with a short-run cap on working capital, one needs a numerical method to compare the options. One can, conceptually, pick a feasible set of options that, in aggregate, enforce the working-capital constraint—an approach akin to invoking a general solver. However, simpler heuristics can be used, such as introducing a time-dependent component for the cost of money.
Dependent options
The rate-of-return metric singles out the best next use of capital, yet the options themselves rarely stand alone. Most tap the same finite pools—cash, shelf space, vehicle capacity—and therefore interact. The interactions can be competitive, when two SKUs vie for the same slot; complementary, when coffee and sugar lift each other’s sales; or sequential, when the second pallet only makes sense after the first has been allocated. Recognizing these couplings is a prerequisite for translating the metric into sound decisions.
A practical heuristic for dependent options is to let the rate of return drive a rolling, one-by-one selection. Pick the option with the highest current rate, update every remaining rate to reflect the resources just committed (or released), pick again, and repeat until all constraints are binding. Computer scientists call this a greedy algorithm; its outcome is a prioritized list. Because each refresh propagates current constraints, the list need not stay in perfect descending order—most rates drift downward as competition stiffens; a few may even rise when a complementarity is unlocked—yet every step remains the best move available at that moment. Where couplings are mild, this greedy pass suffices; where they are tight—vehicle routing or production sequencing—a heavier solver that explores many combinatorial paths outperforms the heuristic while resting on the same valuation bedrock.
For example, consider the challenge of replenishing a retail store. Based on the rate of return, select one unit to add to the store—the one that offers the highest rate. Once this one unit has been, conceptually, added to the store, the rates of return are adjusted. The next unit of the same product has a lower rate of return, as piling up inventory in the same store exhibits diminishing returns. A second unit is then picked based on the revised rates of return. Repeat the process until the store’s shelf capacity is reached.
Greedy optimization offers several major benefits. First, it requires nothing more than establishing the options’ rates of return. Second, the process is simple and observable: the prioritized list provides a complete picture of the algorithm’s progression. Third, when properly implemented, it is typically orders of magnitude faster than non-greedy alternatives.
A naive implementation of this greedy optimization logic incurs quadratic computational cost relative to the number of competing options. For most supply chain situations with thousands of options, quadratic cost is unacceptably high. However, methods—beyond the scope of this discussion—mitigate this overhead, reducing complexity to quasi-linear time.
Most supply chain challenges are well suited to greedy or quasi-greedy methods because the options are loosely coupled. This is no accident: companies deliberately avoid systems that resemble cryptographic puzzles; what is difficult for a computer is also challenging for the human mind.
For instance, a specialized retailer selling three technical products—A, B, and C—typically purchased together in a fixed ratio (e.g., two units of A, five of B, seven of C) to form a kit can simplify inventory management. Instead of tracking three separate items, the retailer treats the preassembled kit $$K$$ as a single inventory item, while maintaining a small surplus of individual components for exceptions. This consolidation reduces complexity and streamlines replenishment.
However, greedy optimization is not suited to every supply-chain scenario—vehicle routing and production scheduling are notable exceptions. More generally, problems involving sequences of highly interdependent decisions do not lend themselves to the greedy method. That said, many—if not most—supply-chain situations can be treated quasi-greedily with minor twists. For example, when placing simultaneous purchase orders to multiple suppliers—each with its own minimum order quantity (MOQ)—a greedy strategy may yield infeasible results, such as a series of incomplete orders that fail to meet thresholds under a limited budget. Yet a minor heuristic suffices to address this limitation. One can iteratively drop suppliers—starting with those yielding the lowest returns—and redistribute the budget among the remaining suppliers.
While the complications arising from dependent options do not undermine the principle of selecting options with the highest rate of return, they may render naive selection methods ineffective. To handle these complexities, dedicated optimization software—often called solvers—is required. Notably, these solvers optimize against a wide array of loss functions (the quantity to be minimized), a topic addressed later in Chapter Decisions.
Uncertain rate of return
Interdependencies were the first hurdle; uncertainty, the second. Even after mapping how options compete for cash, capacity, or shelf space, we still do not know which numbers will materialize when those options unfold. Demand, lead times, yields, and spot prices all waver outside the planner’s control. Yet the guiding principle endures: choose the option with the greatest economic return—measured not at a point, but in expectation. In practice, replace every deterministic payoff with its probability-weighted counterpart: list plausible scenarios, assign each a likelihood, apply the deterministic formula within each branch, then average the results. This produces an expected rate of return that can be compared across options as before.
While conceptually straightforward, the probabilistic perspective presents its own challenges. First, probabilities must be estimated—this is the essence of probabilistic forecasting. Second, the future hinges on decisions not yet made: an option’s rate of return depends on how future options will be selected. Choosing present and future options under uncertainty is the essence of stochastic optimization.
Probabilistic forecasting and stochastic optimization are technical subjects that will be discussed in Chapter The Future. Yet the probabilistic view of the rate of return is no mere theoretical curiosity. With proper software, real-world supply chain options can and should be evaluated by this method. Naturally, approximations are unavoidable—the number of potential futures is infinite—but this does not diminish the approach’s practical value.
Cronyism and regulatory capture
The profit motive is often caricatured as a cold-blooded quest for “more at any cost”. In reality, within a competitive market, profit is simply the applause earned for satisfying the most urgent wants of strangers while spending the fewest scarce resources. Profit is therefore an outcome, not a creed.
Behind the single word “profit” lie three radically different behaviors. Genuine profit seeking earns margins by offering a superior mix of availability, quality, and price. Long-run profit therefore depends on nurturing goodwill with customers, suppliers, and employees. Slash-and-burn tactics that delight today’s spreadsheet but erode tomorrow’s options prove self-defeating once their full rate of return is computed.
A second behavior is the “naked” profit motive popular culture loves to condemn: dumping waste in rivers, bullying suppliers, or bait-and-switching customers. Such conduct is not the logical outcome of profit seeking; it is the outcome of faulty economic calculation. When legal fines, lawsuits, talent flight, and brand damage are brought back to present value, the mirage vanishes. Modern supply chain analytics exists precisely to surface those latent costs by turning them into shadow valuations that discipline every decision.
The third behavior is crony capitalism and regulatory capture. A firm that lobbies for tariffs, exclusive licenses, or mandatory standards does not earn profit; it extracts rent. No customer is served better—resources are merely diverted by political privilege. Rents distort every supply chain signal: optionality shrinks, innovation stalls, and the apparent rate of return is propped up by statute rather than performance.23
The discipline presented in this book targets the first path while inoculating the firm against the latter two. All subsequent techniques assume an open market: they benchmark each option against what a vigilant competitor could do tomorrow, not against a privilege secured in parliament yesterday. Restore genuine competition and the same optimization machinery becomes an engine of wealth creation; let rents accumulate and no amount of forecasting finesse can salvage the economics.
Supra-economic goals
A recurrent objection to profit-seeking firms is the appeal to supra-economic goals—ends that allegedly outrank mere monetary considerations and thus justify overriding the discipline of prices, costs, and opportunity costs. These calls usually come in two flavors. The first is moralistic: a conviction that the firm should advance some ethical or social agenda beyond serving customers. The second is apocalyptic: a prediction of looming catastrophe that demands immediate sacrifice of profitability to avert disaster. Both categories will be examined in the pages that follow.
For companies that operate in open markets, invoking a higher purpose does not dissolve scarcity; it simply relabels trade-offs. Every pallet, man-hour, and coin devoted to one objective is unavoidably withheld from another. Economic calculation is therefore not an optional worldview but the only coherent way to compare alternatives. Institutions that rely on coercion—armies, for example—sit outside the scope of this book precisely because their resources are not allocated through voluntary exchange. For all other organizations, the alleged supremacy of supra-economic goals is an illusion; in practice, they must still decide whether diverting resources from their core mission creates or destroys long-run value.
Moralistic concerns
In supply chain terms, a moralistic concern asserts that people should consume less—or more—of something; hence the company should help steer the desired social change. Details vary by time and place.
For example, in Ruritania, a dominant religion might declare that milk consumption by adults is a sign of depravity, and that milk should be reserved for children. In response, a Ruritanian company might voluntarily reduce both the quantity and variety of milk bottles marketed to adults, thereby proactively enacting the seemingly desirable social change. Yet such a move will scarcely affect the market. Contrary to the popular trope of “economic power”, a company cannot force customers to buy or forgo its products. If the company arbitrarily cuts production of adult-branded milk bottles, a single competitor will raise output to keep demand equally satisfied.
If, however, public sentiment in Ruritania against bottled milk for adults is so strong that many customers boycott other unrelated products, it may be reasonable to divest that line entirely—perhaps by selling the unit if its production can be separated from other lines. This approach differs fundamentally from merely lowering supply in the misguided hope of lowering overall market demand.
Typical moralistic concerns include foods or beverages deemed inappropriate, either for alleged harm to individuals or for harms incurred during production. Veganism is, for example, one of the more recent creeds in an otherwise extremely diverse set of creeds. These concerns also extend to country of origin, often grounded in incorrect mercantilist24 views of trade (e.g., “we are at economic war with Ruritania”) or in perceptions of work conditions in those countries (e.g., “Ruritania lets its companies unduly exploit its citizens”).
Supply chain is concerned with moralistic views only insofar as serving a type of good, or using a particular factor of production, might compromise its general capacity to operate. Brand damage can alienate customers or diminish the company’s appeal to potential employees. These concerns are real and can be quantified just like any other risk. Economic calculations may indicate that such risks are excessive, suggesting that the company should exit a market segment to minimize future losses. However, these calculations should not be mistaken for the personal value judgments of supply chain employees.
What makes the challenge thornier, however, is that employees25 have perverse incentives when it comes to moralistic concerns. By publicly pressing a moralistic concern, in a process referred to as virtue signaling, an employee presents himself as particularly virtuous—more so than his peers. Virtue signaling was and remains effective because corporate politics tend to reward those who conform most to prevailing social norms rather than solely those who drive the company’s profitability.
However, within a supply chain, it is not the virtue-signaling individual who bears the cost of the revised course of action, but the company itself—and, to some extent, the market, which will be less adequately served until a competitor steps in to fill the gap. Thus, corporate executives must remain vigilant against such virtue signaling. Indeed, large companies are especially susceptible to this class of antisocial behavior, which undermines the authority of the nominal hierarchy and its ability to steer the company toward serving customers.
Apocalyptic concerns
Throughout history, a fascination with annihilation has gripped a significant portion of humanity. Exploring these morbid fascinations and understanding why they have endured for millennia is a matter for psychology, sociology, and history. Supply chains are concerned only insofar as these fascinations interfere with the conduct of business. History yields two observations: first, the magnitude and impact of apocalyptic concerns have been vast; second, their specifics have varied dramatically from decade to decade. Let us review a short list of modern apocalypses once in fashion.
-
Overpopulation (18th–19^th^ century): The rapid population growth during the first industrial revolution led numerous intellectuals to believe that population growth would inevitably outstrip food production, leading to widespread famine and societal collapse. (See An Essay on the Principle of Population (1798) by Thomas Malthus.)
-
Degeneracy of the Race (1920s–1930s): Fears of racial and genetic degeneration, influenced by eugenics and pseudoscientific racial theories, predicted the decline of civilizations due to the mixing of races. (See The Passing of the Great Race (1916) by Madison Grant.)
-
Communist Takeover and Red Scare (1950s): The rise of the Soviet Union and the spread of communism led to fears of a global communist takeover, resulting in the Red Scare and McCarthyism in the United States. (See The Naked Communist (1958) by W. Cleon Skousen.)
-
Chemical Agriculture (1960s): The detrimental effects of pesticides, particularly DDT, on the environment and public health highlighted broader concerns about industrial pollution and ecological collapse. (See Silent Spring (1962) by Rachel Carson.)
-
Nuclear Winter and Global Cooling (1970s): Amid the Cold War, fears of a nuclear winter—an aftermath of nuclear war causing global cooling—merged with concerns about a potential natural ice age. (See The Cooling: Has the Next Ice Age Already Begun? (1976) by Lowell Ponte.)
-
Resource Depletion (1970s–1980s): Numerous predictions were made that humanity was running out of essential resources like oil and minerals, leading to societal collapse. (See The Limits to Growth (1972) by Meadows et al.)
-
Access to Fresh Water (1980s–1990s): The depletion and pollution of freshwater resources were predicted to lead to severe shortages and conflicts over water. (See Cadillac Desert: The American West and Its Disappearing Water (1986) by Marc Reisner.)
-
Acid Rain (1990s): Industrial emissions causing acid rain were feared to be devastating forests, lakes, and buildings, leading to widespread ecological damage. (See Acid Rain: Its Causes and its Effects on Inland Waters (1992) by B. J. Mason.)
-
Ozone Depletion (1990s): The depletion of the ozone layer due to refrigerant gases, leading to increased ultraviolet radiation, skin cancer, and ecological disruptions. (See Ozone Diplomacy: New Directions in Safeguarding the Planet (1991) by Richard Elliot Benedick.)
-
Global Warming (2000s): Rising levels of “greenhouse” gases were predicted to cause catastrophic climate changes, including sea level rise, extreme weather, and loss of biodiversity. (See An Inconvenient Truth: The Crisis of Global Warming (2006) by Al Gore.)
-
Climate Change (2010s–2020s): Building on the global warming narrative, climate change encompasses broader concerns about unpredictable and severe environmental impacts, threatening the sustainability of human civilization. (See The Uninhabitable Earth: Life After Warming (2019) by David Wallace-Wells.)
This list spans the past 150 years, though it could extend back millennia. Its purpose is not to validate or refute these threats—such judgments lie outside supply chain and within specialized sciences. Rather, it illustrates the phenomenon’s magnitude and the inconsistency of the threats’ particulars.
Apocalyptic concerns are supra-economic because the view is that, if unaddressed, the world—and thus the market economy—will end or be left in an irretrievably degraded state. By definition, an apocalyptic concern overrides any individual’s particular want. After all, no individual can be anything but insignificant when weighed against the possibility of oblivion for a sizable portion of mankind, and potentially all of it.
For apocalyptic fears that have fallen out of fashion, it is evident that companies which historically allocated resources to support those causes were severely misguided, resulting in significant self-inflicted harm. Few managers would still argue that a hiring policy explicitly discriminating against a minority religion reflects the best economic interests of the company. Similarly, few managers would nowadays argue that generous donations to charities and political movements advocating eugenics genuinely contribute to the betterment of mankind. Yet, not so long ago, both policies were widely practiced by educated and well-intentioned people, operating under what was considered firm “scientific” consensus26.
These apocalyptic concerns disrupt supply chains because their supra-economic claim invalidates economic calculation. Companies, however, still must allocate scarce resources. Apocalypse does not abolish scarcity. There is no alternative to economic calculation when it comes to satisfying customers. If a concern outranks satisfying customers, why allocate anything less than all resources to it?
As King Richard cries in Richard III, “A horse, a horse! My kingdom for a horse!”
More generally, the notion that any concern over allocating scarce resources enjoys supra-economic status, categorically superseding mundane concerns, reflects a profound misunderstanding of economics. There is no alternative to profits and losses.
Non-profit organizations
In most Western countries, tax laws grant special exemptions to organizations styled “non-profit”, unlike those presumed “for-profit”. Whether such tax exemptions and regulatory relief27 benefit society at large lies beyond supply chain. However, as far as general economic understanding goes, this label is very unfortunate and generates immense confusion among the general public. Worse, many authors of supply chain books share the confusion and ascribe quasi-mystical qualities to non-profits, echoing popular but misguided beliefs. These authors mistakenly believe that associations (i.e., so-called “non-profit” organizations) are exempt from the principles of profits and losses. This is patently false. Associations operate under the same economic laws and, insofar as some operate a supply chain, obey the same principles. Consequently, a technique meant to improve supply chains is equally valid—or equally invalid—regardless of taxation status.
A few facts suffice. Both kinds of organizations are driven by entrepreneurs. Some entrepreneurs who run for-profits fail and end poorer than they started. Conversely, some who run non-profits succeed and command salaries far exceeding the national average. Some profitable companies distribute dividends, while others do not. Both companies and associations see corruption scandals. Both require customers to fund them28. Customers have expectations, and in both cases, if a competing organization delivers more for less, customers favor the more efficient organization. Both expand when they collect more than they spend; both go bankrupt when they spend more than they collect.
Calling an association “non-profit” is a misnomer: profit and loss are the basic facts of economics. The organization collects and spends resources. If collections exceed outlays, there is profit. If outlays exceed collections, there is loss. Profit and loss exist even without an accountant. When an accountant keeps a ledger, the situation is clarified; he creates neither profit nor loss, he merely reveals them.
In the market economy, all actors strive to address customers’ most pressing wants. This is true for companies and associations alike. For example, if customers are concerned with deforestation, they may spend to plant trees, or have them planted. By giving to an association dedicated to reforestation, customers make a value judgment: they buy future forests; a world with more forests is their reward. If presented with several equally dedicated organizations, and one can grow twice as much forest per coin, customers will favor it. If customers later discover that the association spent all it collected on advertising and planted no trees, they will rightly feel defrauded and likely abstain from further donations.
The absence of a direct correspondence between a customer’s payment and the service rendered by an association is merely a technicality. The same pattern appears in business: urban waste collection is usually paid via an annual lump sum rather than per service rendered, yet many such providers are for-profits.
Calling an organization “non-profit” is as nonsensical as calling it “gravity-free”. Natural laws admit no objection; those that govern economics are no exception. How profits are allocated—whether distributed to shareholders, reinvested in further capital goods, or used to raise employee wages—is a technical detail. Taxation schemes materially affect these technicalities, but not in a way categorically different from, say, subsidizing a sector such as solar energy. It is merely one of many approaches by which governments partition the market in distributing subsidies. Subsidies confer no special economic power on associations.
Ultimately, entrepreneurs’ inner motivations remain unknowable to external observers, whether their organizations are for-profit or non-profit. They may be genuinely moved by the improvement of their fellow men, or merely chase vanity and the publicity that attends a successful association. Men can lie; they can also change. What starts in selfish intent may, over time, become genuine altruism; the converse is also possible.
Thus, supply chain cannot categorically distinguish alleged non-profits from for-profits. This distinction belongs to tax codes, not to supply chain. Authors who claim that associations possess unique supply chain properties are either engaging in virtue-signaling—posturing with noble causes—or are misled by ideology, attributing quasi-mystical qualities to these organizations.
Armed forces
Non-profit organizations illustrate how profit and loss remain inescapable even under altruistic pretenses. Armed forces offer a distinct case: an institution squarely within economics yet lacking the usual for-profit feedback loops. Armies do not earn revenue by selling services to a clientele; rather, they are funded through taxation raised by a sovereign power. This reliance on state coercion does not exempt them from scarcity; it only changes how resources are allocated. Economic laws do not evaporate in the presence of arms and uniforms: every bullet and barrel of fuel meets the same constraints that govern the rest of the economy.
From an economic perspective, a “military” is among the most extensive central-planning structures in modern societies. The armed forces must maintain equipment, procure supplies, support troops, and project power across various geographies. However, these activities cannot be assessed using the profit-and-loss framework applied to market actors. While a private company constantly balances expenditures against customer payments, the military is funded without voluntary exchange, and thus lacks direct market discipline. The invisible hand that fosters efficiency in private ventures is replaced by top-down directives meant to integrate strategy rather than short-run financial returns.
Yet the absence of profit signals does not spare the armed forces from the problems of economic calculation. As Ludwig von Mises demonstrated in Socialism (1922) and Human Action (1949), when a central planner attempts to guide resource allocation outside the pricing process, there is no reliable gauge to separate profitable from unprofitable uses. Without the discipline enforced by free markets, misallocation is almost guaranteed. Such misallocations remain hidden in peacetime—when the budget appears as a single line item—but become evident when the armed forces face shortages of critical gear or surpluses of obsolete equipment. The same challenges afflict large companies running complex supply chains, but armies face them on a grander scale, with life-or-death consequences.
In modern times, defense budgets running into the billions reflect complex trade-offs made at the highest political level. The classical “guns versus butter” analogy captures how each unit devoted to the military is one not spent on civilian goods—schools, hospitals, roads, or other productive ventures. The armed forces do not “profit” when they fulfill their mission; rather, they receive continued political support to do more of the same. When a new generation of fighter jets or warships is commissioned, there is no queue of customers whose willingness to pay would justify the spending. Instead, the justification emerges from strategic imperatives, real or perceived, that fall outside the marketplace’s regular accountability mechanisms.
Naturally, militaries try to mitigate the pitfalls of central planning by resorting to surrogate metrics: “readiness”, “mission capability”, “force projection”, and other strategic or bureaucratic indicators. While useful, these indicators are no substitute for the immediacy of profit and loss. A victorious campaign produces no revenue surplus for the armed forces, just as a protracted conflict does not book directly as a loss. Extended conflicts can yield larger defense budgets and expanded infrastructure, absent any clear demonstration of effective economic use. Decision-making reverts to politics—what the public or its representatives deem justifiable—rather than to trade-offs shaped by voluntary demand.
These constraints generate unique supply chain challenges for the armed forces. Demand is highly uncertain and driven by geopolitical events rather than consumer preferences. Large stockpiles are maintained to hedge against crises, and lead times can be extraordinarily long for specialized equipment such as nuclear submarines or stealth aircraft. When an army lacks sufficient capacity to produce critical components itself, it must turn to private defense contractors. At that juncture, a portion of the supply chain reenters the realm of market competition, at least on paper: suppliers are paid according to negotiated prices, and nations often try to pit multiple contractors against each other to drive costs down. However, political interference and regulatory complexities often distort these processes, leading to “cost-plus” contracts and other arrangements that too often degenerate into crony capitalism. As a result, while a veneer of market discipline can exist, it is usually partial and overshadowed by bureaucratic politics.
In the same vein, Hayek’s knowledge problem in The Use of Knowledge in Society (1945) applies to the military as well. Information relevant to resource allocation is widely dispersed. Local units have up-to-date insights into their logistics and strategic constraints, but higher-level commands make the grand allocation decisions. Strict top-down hierarchies suppress bottom-up signals, depriving upper echelons of the local knowledge needed to make informed choices about procurement, deployment, or maintenance. De facto, the arms industry acts as a closed microcosm of central planning: no matter how competent or well-intentioned the officers and bureaucrats, critical information is lost in transit, heightening the risk of misallocation.
Nonetheless, all relevant resources remain scarce. Missing supplies can sink entire campaigns, and, conversely, excessive stockpiles of obsolete gear can hamper the military’s ability to finance modernizations. Moreover, as with any large-scale bureaucracy, routine improvement efforts add layers of process: new committees, new oversight boards, new planning steps. Over time, these layers develop inertia of their own, making course corrections harder. Without a profit test to cull waste, entire procurement lines may survive long past their useful life. Politically favored programs persist even when they yield little advantage on the battlefield. The result, economically, is structural inability to optimize resource use. Militaries are not free to shed these constraints: they cannot become agile forces guided by profit. They fulfill functions outside voluntary collaboration, and the price of the arrangement is the absence of direct market feedback.
A final peculiarity of the armed forces is that they are subject to abrupt reorientation, typically following changes in political leadership. Rebudgeting can be drastic, upward or downward, depending on the perceived threat. Shifting alliances, new treaties, or war-weariness can all reshuffle military priorities. While private businesses in a free market also adjust to sudden shifts, they can rely on profit signals to guide them toward better allocations. By contrast, armies must redesign their supply chains according to top-down instructions, incurring the overhead of a large bureaucracy that must simultaneously produce new doctrines, scrap old ones, and realign entire fleets of vehicles and equipment.
Thus, armed forces exemplify an extreme manifestation of the same planning conundrums faced by large integrated enterprises. The state does not measure the success of its armies by how effectively they satisfy consumers’ wants; no direct channel lets consumer dissatisfaction act on the military budget. Yet resources remain scarce, and the need to distribute them persists, so the logic of economics still applies. The absence of profit-and-loss signals can be partly offset by surrogate metrics, but such substitutes offer nothing close to the breadth and refinement provided by market prices. The upshot is that, even if armies operate in a non-market context, every conflict, every peacetime mobilization, and every behind-the-scenes logistical preparation exists under the iron law of scarcity. While the state of war is often treated as pure politics or strategy, at its heart it is also an elaborate allocation problem no decree can escape.
Solutions and trade-offs
There are no solutions. There are only trade-offs.
A Conflict of Visions (1987), Thomas Sowell.
Supply chains are strictly bound by the laws of economics. Each allocation of a resource forecloses all alternative uses that were readily available. A decision is correct only if there is no better use for the resource; profitability is insufficient—it must be maximal.
Any inventory allocation carries the opportunity cost of the best alternative. In a retail network, once a warehouse unit is dispatched to a store, it is no longer available to serve other stores. If it was the last unit and it goes to a slow store while others sell briskly, that allocation was likely a mistake. Better to route the last unit to the store with the most urgent demand. Either way, the unit will likely sell. Yet sending it where demand peaks aligns the network with the customers’ most urgent wants. If the unit at the mediocre store lingers, it may be returned to the warehouse and redispatched to a more suitable store. However, this incurs extra transportation costs, and while in transit the unit serves no one.
Selling now trades off against the chance to sell more profitably later. The later, more profitable outcome may not require a higher price. For example, a screw bought at 1 coin and sold at 2 yields 1 coin of gross margin. However, if this screw—the last one in inventory—belongs to a bundle (say, a machine sold for 100 coins with component costs of 50), then selling the screw alone yields 1 coin now but forfeits 50 coins should a bundle buyer arrive before more screws can be secured. If the firm expects a prompt bundle buyer, selling the screw alone is the wrong supply chain decision despite being profitable. The correct move is to keep the screw in stock to preserve the bundle sale.
Resources are varied, and the resulting trade-offs are even more so. Yet many supply chain authors reduce decisions to inventory costs vs stockouts, cost vs cash vs service, or supply vs demand. These low-dimensional perspectives are misguided. By elevating a few engineering trade-offs—rooted more in bookkeeping than in economics—to first principles, they obscure the only universal yardstick: economic calculation.
Across any industry, supply chain is a continuous walk along a tightrope of mutually exclusive improvements. Drop the ticket price and volumes usually spike, yet every percentage point conceded bites straight into the gross margin—and, in luxury segments, can cheapen the brand so severely that unsold units end up destroyed rather than discounted. Broaden the assortment with extra colors or sizes and customers applaud, but procurement, storage, and quality‐control costs climb just as fast. Add backup suppliers to hedge late deliveries, and the resilience is bought with weaker purchasing leverage. Keep stores open longer and extra footfall appears—along with a fatter wage bill. Hire temporary pickers to clear a seasonal spike and throughput improves while errors and shrinkage rise. Switch to a boutique component maker and the finished good feels premium—provided the small plant does not miss its slot. Relax the return policy to nurture loyalty and reverse-logistics expense balloons. Rotate perishables daily to maximize freshness and the shelf looks immaculate, yet deliberately thinner inventory guarantees more stockouts. Every lever carries this dual character; planners juggle dozens of such dilemmas at any moment.
This list could be extended almost endlessly across every vertical and its specific trade-offs. Moreover, nearly every named resource can be further decomposed. The label is handy but ultimately an approximation. For example, let’s say that the warehouse capacity is 10,000 $$m^2$$. In reality, not all of those $$m^2$$ are equal. Some sit near loading docks or conveyor endpoints. Capacity may be refined to 2,000 “premium” $$m^2$$ and 8,000 “regular” $$m^2$$.
One can always impose an arbitrary taxonomy on resources and trade-offs with as many classes as one fancies. Within a given company, such a taxonomy may ease communication. Across the market, such exercises are largely pointless. It yields no better results than the common tongue.
The only practical approach to these trade-offs is economic calculation. Economic calculation balances these contradictions to reach an adequate supply chain decision. It also makes clear that no decision is an absolute good—a pure source of profit—but a mix of expected gains and losses.
Valuation concerns
Supply-chain decisions are, at bottom, wagers: each one stakes a scarce resource today for a stream of uncertain consequences tomorrow. Trying to evaluate all those wagers in a single, monolithic calculation is hopeless—the decision space is far too large and changes too quickly. The practical remedy is to slice the grand calculation into bite-sized fragments—one fragment per option—and to equip every fragment with a valuation. A valuation is not an accounting curiosity but an instrument: a money-denominated score that renders options commensurable, so they can be compared, ranked, and picked piece-by-piece rather than all at once. Good valuations expose opportunity costs and power the optimization engines introduced earlier; flawed ones quietly steer the firm toward systematic misallocation.
Seen in this light, valuations form the indispensable bridge between the concrete world of pallets and trucks and the abstract realm of economic calculation. They translate heterogeneous events—an extra day of shelf life, an irritated customer, a late vendor payment, a congested pick aisle—into one scale that software can manipulate at millisecond pace. Building this bridge always follows the same route.
First, the company frames an explicit economic model that makes clear which flows matter and how they connect to profit and loss. Second, it attributes every observed coin to the decisions that generated it, so each micro-choice bears its fair share of shared costs and benefits. Third, it projects those cash flows forward, embracing uncertainty and adding shadow adjustments where a bare ledger would miss material risks or benefits such as goodwill or brand damage.
The next subsections examine these three steps in turn. They show how an explicit economic model anchors the calculation, why attribution is both unavoidable and contentious, and how projections—augmented by shadow and private valuations—turn yesterday’s data into tomorrow’s actionable insight.
The economic model
The economic calculation starts with an explicit economic model: a concise yet transparent mapping that links every operational choice—buy, make, move, hold, price—to its projected financial footprint. In practice, the model translates the flow of atoms into flows of coins by specifying four elements: the unit of analysis (SKU, pallet, machine-hour, etc.), the time horizon over which consequences unfold, the attribution rules that assign shared costs and revenues, and the discounting convention that reconciles different dates. Its purpose is not prophetic accuracy; it is to surface the otherwise invisible trade-offs so that mutually exclusive options can be ranked on a single yardstick—risk-adjusted return. Because supply chains operate on several intertwined time-scales, multiple sub-models usually coexist: a short-cycle replenishment model centered on inventory rotations, a procurement model encoding supplier rebates and MOQs, a capacity model showing how production constraints convert into cash. Each is a facet of the same discipline: turning physical decisions into discounted cash flows.
For example, one of the simplest economic models is that of the pure trader. The pure trader merely buys and sells goods without transporting them, often leaving custody with the seller for an agreed period. The strategy of the trader consists of buying low and selling high. The relevant model reduces to projected gross margin: per-unit sell price minus per-unit buy price. Variability comes from the sell price, unknown at purchase time. Pure traders serve various functions in the market, such as protecting sellers against price variability or sparing them interactions with numerous smaller clients.
While the pure trader is an extreme simplification, real-world supply chains invariably entail numerous subtleties. The economic model is always approximate. Some approximations are intentional, as they help to keep the model manageable. For example, all goods are perishable: given enough time, even so-called stainless steel rusts. Yet for many goods, shelf-life is so long that perishability is negligible. In such cases, omitting perishability from the model is reasonable.
Other approximations are unintentional, reflecting operational uncertainty. Knowledge can be refined, but it will remain incomplete. For example, stock may harbor defects noticed only by customers, some of whom later request refunds or exchanges. Quality control may push defects to near-negligible levels, yet the possibility remains and knowledge never becomes absolute.
Designing the model—including its structure and intentional approximations—is a problem of general intelligence. Those problems do not lend themselves to “mechanistic” resolutions. The adequacy of the model is always a matter of high-level judgment. One can empirically falsify an economic model—or at least quantify its approximations and judge them excessive—but designing the model itself requires higher forms of reason. Guidance and principles can be provided, but no concise formal rulebook will accomplish this exercise satisfactorily.
Yet while the model’s design remains a matter of general intelligence, a great deal of automation can still be accomplished. The model’s evaluation and its use in decision-making can be fully automated. The model must be revised as the company and the market evolve; however, this process is usually slow compared with the fluctuations in the flow of goods and materials. For most companies, a relevant economic model lasts months if not years; a flow datum—such as a stock level—lasts hours.
Implicit valuations
Many companies sidestep the problem of economic calculation for supply chain by resorting to percentage-based policies and similar noneconomic methods. Those policies govern decision-making in supply chain. They typically mix manual and automated operations without explicitly adopting an economic perspective. The sole redeeming quality of this approach is avoiding the overhead associated with economic calculation itself. For the rest, no matter how much success the company enjoys from factors independent of its supply chain—such as superior branding—this approach can no longer be considered anything but backward.
For a microbusiness—a single grocery, a family-run workshop, a corner repair shop—the absence of formal valuations is not backward; it is pragmatic. The owner’s scarcest resource is personal time: every hour spent assembling a spreadsheet competes with wiping down shelves, talking to customers, or simply keeping the lights on. When two hours of arithmetic can at best trim a handful of coins from next week’s purchase order, those hours deliver a higher, safer return when invested in upkeep or direct selling. In that setting, simple rules of thumb—“re-order when only two units remain”, “never let the bread rack go empty”—approximate the optimum well enough. Once volumes, products, and locations grow, the arithmetic flips: the cost of misallocation quickly outstrips clerical overhead. Beyond that threshold, clinging to implicit valuations is no longer frugality; it is self-sabotage.
With this caveat settled, let us return to the main thread of our argument. Safety stocks, min/max reorder points, and time-based buffers are archetypal noneconomic policies. Each spits out a number—percentage, unit count, or “days of demand”—that triggers an order, yet none carries a valuation expressed in coin. They are calibrated to hit engineering targets—fill rate, cycle time, visual comfort—without ever asking what the next unit is worth. Capital cost, margin variance, and opportunity cost are nowhere in the formulas. A company can execute such policies flawlessly and still misallocate resources, because the yardstick they optimize is disconnected from profit and loss.
However much these policies ignore economic calculation, profits and losses still occur; the calculation can be evaded—its consequences cannot. For example, there is a common misconception that percentages should be steered toward 100%, as with service levels steering safety stocks. There are obvious diminishing returns to increasing quantities held in stock. At some point, write-off costs from excess inventory equal stockout costs; beyond that, further increases in stock levels generate losses. When noneconomic policies are used, there is no reason to believe their internal parameters are close to the settings that would maximize returns.
As a silver lining, those settings are unlikely to be catastrophic: if they were, the company would have ceased operations. Survivorship bias looms large: only companies that avoided bankruptcy remain observable; the rest, by definition, are gone. Thus, to some extent, a noneconomic policy that has kept a company solvent over the years demonstrates some merit.
Conceptually, every noneconomic policy can be rewritten as an economic one. Introduce a series of valuations chosen not for their adequacy to real profits and losses, but for their capacity to numerically yield the same decisions. This exercise clarifies the fundamental defect of noneconomic policies: they are economic policies built on flawed valuations.
Thus, any noneconomic policy can be improved by first rewriting it explicitly as an economic policy, then bringing the valuations closer to their real values. Such a roundabout course is mostly of theoretical interest; in practice, there is rarely any justification for addressing the situation so convolutedly. It may occasionally be warranted if the legacy transactional system is inflexible and can operate only on safety stocks, for example. In that case, if the company will not or cannot upgrade its system, it is forced into the tedious process of making implicit valuations explicit and readjusting them to more sensible values. This process involves substantial reverse engineering, is prone to implementation errors, inflates maintenance costs, and must remain an option of last resort.
With noneconomic flow policies now reframed as crude—often faulty—valuations, two practical hurdles still stand between bookkeeping and actionable micro-decisions. Attribution asks: “Which past coins belong to which past choices?”—it untangles the causal threads that weave costs and revenues together. Projection then asks: “Given those threads, what pattern of coins will tomorrow most likely bring?”—it pushes the picture forward under uncertainty. The next two subsections tackle these twin challenges in that order.
Spending and revenue attribution
Attribution reconnects every coin that flows through the ledger to the operational choice that precipitated it. In the world of the pure trader, the mapping is trivial: one purchase, one sale, one gross margin. The moment a company stores, routes, transforms, or promotes goods, this one-to-one symmetry vanishes. A single inbound container replenishes hundreds of SKUs; one markdown lifts a whole category; a missing gasket halts an entire assembly line. Unless those braided effects are disentangled, the “value” of shipping one extra pallet or shaving a day off lead time is anyone’s guess. Attribution therefore provides the quantitative causality that turns raw ledger records into decision-ready data and makes a microanalytical, option-by-option evaluation possible.
Let us consider the problem of establishing the per-unit purchase price for a series of products that can all be acquired from the same supplier. This per-unit purchase price is critical for deciding whether units should be acquired from this supplier or from a competitor. In our example, the supplier puts unit prices on display; however, those tag prices come with strings attached. Indeed, the supplier charges a flat delivery fee, reflecting the cost of dispatching a truck. The client company faces a trade-off: small orders with low inventory costs and high delivery fees; large orders with high inventory costs and low delivery fees. Whatever the size and frequency of purchase orders, the supplier’s invoices include a mix of item lines and delivery lines. The attribution problem consists of deciding how those lines should be allocated to the calculation of per-unit purchase prices.
For clarity, assume the client orders two products: one in large volume and the other in small volume. In practice, the only “economical” way to order the second product is to piggyback on an order that includes many units of the first product. The second product “tags along” in the order stream with the first one. In this situation, it is tempting to allocate delivery fees to both products in proportion to spend. Thus, nearly all the delivery fees would be allocated to the first product, leaving only a small portion for the second. Based on this attribution, both products will see their “adjusted” purchase price inflated by a fixed percentage. However, this line of reasoning ignores the coupling of the two products. If the company switches the first product to a cheaper supplier, it might end up worse off, because savings on the first product are offset by delivery fees still incurred to order the second product. Such a decision effectively inflates the per-unit price of the second product.
The attribution problem is not, however, restricted to spending; it also impacts revenues. For example, let us consider a store offering a substantial discount on a given product. Revenue for this product spikes as more units are sold, albeit with a lower gross margin; the promotional effect does not stop there. Customers also purchase complementary products, prompted by their initial interest in the discounted product. Thus, revenues for several complementary products rise as well. It may appear that these revenues were generated at full price, but this perspective is misleading. Without the discount on the “flagship” product, those other products would not have been purchased to the same extent. Thus, instead of naively assigning the entire revenue contraction—caused by the discount—to the first product, it is more correct to spread this contraction across all complementary products in proportion to their indirect promotion.
More generally, in supply chain, couplings are ubiquitous: items are jointly purchased, transported, stored, transformed, and sold. Each of these operations involves fixed or shared costs. Conversely, nearly all mechanisms generating revenues come with spillover effects: brand awareness and loyalty, impulse buying beyond the original intent, etc. Attribution reorganizes spending and revenues in ways more closely aligned with decision-making. As such, attribution is one of the key mechanisms to perform the economic calculations needed to identify the best options available to the company. To a large degree, attribution defines the very structure of the economic calculation.
Attribution cannot be mathematically proven correct; it can only be shown incorrect if it conjures spending or revenues out of thin air, or if the underlying arithmetic is faulty. However, those issues are relatively trivial, and even flawed attributions may still meet such a basic standard of “correctness”.
In most countries, there exist a great many rules, usually of the accounting variety, to perform such attributions. For example, those rules may aim to prevent “selling at a loss”—according to some semi-arbitrary definition of “losses”29. Those rules are intended to fulfill taxation or regulatory purposes. Their validity does not extend any further than their original goal. It would be a grave mistake to recycle those rules for analytical purposes. Supply chain must abide by regulatory requirements, but that is all.
Every attribution scheme comes with its own distortions. From an informational perspective, attribution is a dimension-reduction technique. Back to the high-volume/low-volume example above, we have two item prices and one delivery fee. By attributing the delivery fee to the products themselves, we project a three-dimensional situation into a two-dimensional space. This projection simplifies the decision-making process, as each product gets a shadow price reflecting the delivery fee. However, it may lead to misguided decisions if those shadow prices misrepresent the original situation. In general, folding dimensions cannot perfectly preserve all properties of the original space.
Optimization techniques capable of operating in high dimensions may alleviate, or even eliminate, the attribution problem. For example, if we have a numerical recipe capable of directly handling the six dimensions of a pair of suppliers competing for the same two products (i.e., four per-unit prices and two delivery fees), then no attribution is needed: the recipe embraces the higher dimensionality of the problem.
Attribution belongs to economics rather than economic history, in that it takes the accuracy and completeness of records as given. It simplifies the problem space—through dimension reduction—while maintaining a relevant economic model, i.e. one that reliably represents spending and revenue baselines for valuing options. Once those baselines are established, one must proceed to the second mechanism: projection.
Projected spending and revenues
Economic history must not be confused with economics proper. While past spending and revenues can be made certain—aside from occasional clerical errors—future spending and revenues are always uncertain. Their uncertainty is irreducible. No matter how recent or numerous the transactional observations, they can never tell the complete story of what will come to pass, because the future depends on decisions that have yet to be made. Some of those decisions will be made by the company itself; others will be made by third parties—customers and suppliers, for example.
As supply chain options are picked for their assessed economic returns, they ultimately rely on projected (and therefore inherently imperfect) valuations. There are situations where valuation discrepancies between the recent past and the near future are deemed insignificant, but those situations are rare. The opposite dominates supply chain. Let us briefly illustrate this proposition.
When a retailer decides to add one more unit—tagged at one coin—to the shelf, two complementary projections must be made.
First come the projected outflows. From the instant the item enters the store until the day it leaves, it ties up capital and absorbs carrying costs: rent for the floor space it occupies, wages for staff who move or count it, utilities, shrinkage, insurance, and the opportunity cost of cash immobilized in the unit itself. These expenses accrue day after day, and their expected value is strictly positive; were they truly negligible, the rational strategy would be to expand the sales area without bound—a reductio ad absurdum that confirms the costs are real and material.
Second come the projected inflows. The one-coin label is only a candidate price: if demand lags, the unit may have to be discounted; if it degrades, it may be written off; if competitors undercut, the store may follow; and, in the opposite direction, a sudden demand surge or general inflation can push the realized price above the tag. Thus the expected cash recovered at sale is almost never exactly one coin—most often lower, occasionally higher. Only after both streams are discounted to present value can the retailer decide whether stocking that extra unit raises or lowers long-run profit.
More generally, the projection problem is as thorny as the attribution one. While the example above is straightforward—the measurement is merely delayed by one inventory cycle—many concerns are more elusive empirically. Yet, those concerns can also prove to be critically important. There is no correlation between ease of measurement and criticality for the business.
Let us consider a fashion brand discounting at season’s end to liquidate the current collection. There are two distinct forces at play. First, the discount increases the sales volume, but at the expense of the gross margin. Second, the discount trains customers to expect future comparable discounts, eroding willingness to pay: some will delay purchases until the sales periods. This erosion might slow; yet, given enough time, its effect on the brand’s revenues is considerable. It may well dwarf the first one for many, if not most, brands. Yet, while quantifying such erosion is not impossible, it is far from being straightforward.
Every valuation involves a projection—a numerical statement concerning the future, a “forecast” not restricted to demand alone. Historical spending, revenues, and their attributions provide the baselines on which to compute these projections. However, historical baselines should not be equated with their projections. The future is contingent on decisions that have not been made. Any projection that appears accurate can become completely inaccurate if subsequent decisions are revised. This problem is fundamental and cannot be eliminated by statistical methods alone, however sophisticated; it can be mitigated by methods that acknowledge inherent uncertainty.
The economic performance of a supply chain decision largely depends on the quality and reliability of its underlying projections. While addressing this challenge typically involves a great deal of statistics, it is an error to treat it as a purely statistical problem. Indeed, a projection must, first and foremost, be assessed on its economic relevance—a criterion more demanding than mere statistical accuracy.
The remainder of this section zooms in on two special categories of projections that deserve dedicated treatment. Shadow valuations come first: they are numbers that do not appear in any ledger line yet must be conjured into existence to capture elusive but economically decisive effects. We then turn to private valuations, the internal transfer prices applied to intermediate components that never face the outside market yet still compete for cash, space, and attention inside the firm. Both families follow the general logic just established, yet each brings its own pitfalls and practical techniques.
Shadow valuations
The economic impact of a decision is usually tallied by projecting the cash spent today and the cash earned tomorrow. Yet many second-order effects—lost goodwill, supplier distrust, employee fatigue—never hit the ledger, even though they later translate into hard cash. Ignoring them yields a caricature of a short-sighted, spreadsheet-driven accountant who “optimizes” this quarter while quietly destroying next year. To escape that trap, practitioners must introduce shadow valuations—explicit, money-denominated estimates that render invisible consequences tangible. These figures are assumptions—there is no historical baseline for them—but without such adjustments, the calculation collapses into naive cost-cutting that sacrifices long-term optionality and, ultimately, profit.
Consider a retail example: if a customer does not find what he expected in a store, his purchase is at least delayed—and may even be canceled if he loses interest. He may instead visit a competitor, find the product, and eventually prefer that store. Such service failures occur silently, leaving no trace in transaction records, yet they have real consequences—both short- and long-term revenues suffer.
If a retailer values its inventory solely on observed sales, it overlooks how inadequate service erodes its customer base. In other words, by undervaluing inventory—and setting aside shelf-capacity constraints—the store ends up holding suboptimal stock. Recognizing this, the company may introduce a “stockout penalty”, a shadow valuation designed to capture the long-term cost of lost sales. When incorporated into the economic calculation, this penalty raises the perceived value of holding inventory, encouraging the store to carry more.
More generally, the goodwill30 that a company enjoys with its customers, suppliers, and employees is best captured through shadow valuations. Although specifics vary across organizations, most supply chain decisions affect goodwill. Therefore, incorporating shadow valuations better accounts for these broader consequences when evaluating options.
For example, suppliers may be harmed by erratic purchasing—fluctuating order sizes and inconsistent frequencies. By smoothing order flow and making demand more predictable, companies encourage suppliers to lower their prices. In this context, a shadow valuation that penalizes order variability steers purchasing behavior in the desired direction.
Similarly, employees suffer under excessive workloads. For instance, if a grocery store’s staff receives more items in a day than can be shelved during regular hours, they may be forced into unplanned overtime. Moreover, inventory left scattered rather than properly arranged frustrates both employees and customers. A shadow valuation that penalizes excessive delivery volumes helps smooth warehouse dispatches.
In practice, shadow valuations are often rough estimates—and that is acceptable; being approximately right is preferable to being exactly wrong. Overlooking elusive yet consequential factors is misguided. This oversight leads to “idiot savant” solutions, which may exhibit great technical sophistication while still missing the point. Although we will later discuss how experimental optimization offers a more robust solution for valuations, for now informed approximation is preferable to ignoring significant factors.
Private valuations
A fundamental feature of the market economy is the establishment of prices. For finished or “sellable” goods, companies must operate within the constraints set by market prices. A company can steer its prices downward—by innovations that slash costs—or upward—through careful branding, as with luxury goods. However, market prices always apply. Even a company with a patent-backed monopoly must adjust its prices with regard to alternatives in the market. Market prices do not require perfect substitutes.
In contrast, non-sellable goods do not directly benefit from market prices. Such goods are common in manufacturing, where production often requires numerous intermediate components—some assembled from subcomponents sourced externally. These intermediates typically exist only within the company’s production process and lack the complements (technical documentation, established sales channels, marketing materials, compliance procedures) that would render them sellable. However, not all intermediate goods are non-sellable; for instance, a manufacturer may produce both pumps and air-conditioning units, where the former is a part of the latter yet is also sold independently.
Non-sellable goods require investments in both production capacity and sufficient inventory. When a non-sellable component is used exclusively in one finished product, its valuation is relatively straightforward31—the cost of the finished product can be allocated proportionally across its components. However, when a component serves multiple finished products, valuing each additional unit in stock becomes more challenging, as it depends on the probability of its use across products.
For example, suppose a manufacturer uses a component in two products sold in equal volumes: Product A yields a 10-coin gross margin and Product B only 1 coin. The value of producing an extra unit of this component depends on which product it supports. If ample stock of Product A exists while Product B is out of stock, the next unit is more valuable for Product B. Conversely, if Product A is out of stock and Product B is well supplied, the additional unit is more valuable for Product A, which offers a higher margin. Thus, the extra component’s value depends critically on downstream inventory levels and product margins.
In practice, a manufacturer may handle numerous products and components. Production cycles often differ, and demand fluctuates in both volume and price. Each product typically maintains its own stock, often distributed across multiple locations, and production may occur in batches of varying size. These conditions complicate the valuation of component stocks.
The rate of return of any investment related to non-sellable goods—whether acquiring capital equipment or using input resources to produce extra units—must be evaluated by its projected impact on the company’s revenue and expenses. Like the attribution introduced previously, private valuations serve as a dimension-reduction technique, summarizing complex, multidimensional effects into an expected return.
In larger, vertically integrated companies, private valuations grow in prevalence and importance. In such firms, non-sellable goods often outnumber finished products and account for a higher share of inventory value. Consequently, valuing these assets becomes critical.
The role of the market
Private valuations are inherently difficult. As companies grow larger and more vertically integrated, their supply chain issues increasingly resemble the resource-allocation problems governments face under socialism and its central planning. Indeed, socialism has been implemented many times, in many places; as leading economists predicted in the first half of the 20^th^ century, all those attempts failed and invariably led to general impoverishment. Socialism, with its corollary of central planning, fails because it tries to solve the problem of valuing scarce resources—a challenge that only the free market can resolve at scale. Similarly, although companies can perform accurate private valuations at small scale, large firms often encounter significant diseconomies of scale—flawed private valuations being a key factor—that limit growth.
In a free market, prices signal to economic actors how resources should be allocated. When demand exceeds supply, prices rise, prompting investments in the scarce product. Conversely, an oversupply drives prices down, encouraging a shift toward more profitable opportunities. When a supply chain is relatively simple—with few products, minimal intermediate goods, and short lead times—translating market prices into resource-allocation decisions is straightforward. In these cases, market prices themselves resolve a critical piece of the supply chain challenge without the need for deep insights or advanced analytics32.
As large companies expand their supply chains in depth and complexity, the link between planning and market-price feedback weakens, and challenges multiply. This issue was well known to the entrepreneurs of the 19^th^ century, but in his Testament of a Furniture Dealer (1976), the founder of IKEA, Ingvar Kamprad, puts it plainly, stating:
But do not forget that exaggerated planning is the most common cause of corporate death.
Indeed, much like governments attempting socialism, large companies have no alternative but to appoint a caste of experts—such as supply and demand planners, category managers, demand forecasters, and inventory managers—to address the allocation of resources. Although these roles differ only slightly in scope, the overall challenge is immense: companies must account for hundreds of economic factors and establish thousands of private valuations, all while facing inherent future uncertainties. Flawed valuations beget unprofitable decisions.
As an organization scales, incentives shift: middle managers increasingly prefer avoiding blame to chasing incremental gains, and risk appetite shrinks with firm size. Consequently, they rely on established processes, often adhering to them rigidly regardless of economic merit. When these processes repeatedly yield unprofitable outcomes, the bureaucracy responds by adding more rules, extending checklists, and incorporating additional workflow steps. In essence, the bureaucracy expands regardless of economic justification.
Diseconomies of scale—not a shortage of ambition—usually put the brakes on very large, vertically integrated firms. As organizational layers multiply, private valuations drift, and every new tier magnifies the resulting misallocations. To regain agility, companies typically shed peripheral activities—through outsourcing, spin-offs, or joint ventures—and redouble their efforts on segments where their comparative advantage is obvious and where external market prices offer a rapid, unambiguous feedback loop.
Information
Every supply-chain decision is a bet under uncertainty; to be any good, it must be informed. Somewhat unexpectedly for the lay reader, “informed” is not merely a metaphor: since the 1940s, information theory has given the word a precise, quantitative meaning. This chapter clarifies what “being informed” entails for a modern supply chain.
Two facts motivate what follows. First, contemporary supply chains are irreducibly digital objects. Their states and histories live inside business systems; calling a supply chain “digital” is redundant, because no sizable chain can operate without software. Second, mainstream supply-chain theory routinely collapses data into information: it treats the quantities it aims to decide upon—demand, lead times, service levels—as directly observable. In reality, the firm only holds crude traces (sales orders, shipments, returns, stock corrections) from which those quantities must be inferred. Ignoring this distinction is precisely why elegant formulas so often fail when confronted with the raw exhaust of enterprise systems. This mismatch between theory and practice drives many failed initiatives: firms try to graft pre-computer ideas onto post-computer infrastructures.
To navigate this landscape, we must separate three notions that everyday language blends: data, information, and knowledge. Data are the recorded symbols—bytes in tables, log lines, barcodes. Information is the capacity of those symbols to resolve uncertainty, a quantity Shannon taught us how to measure. Knowledge is the causal structure that lets a mind—human or machine—turn information into decisions that improve the firm’s returns. Confusing these notions breeds naive recipes that look sophisticated on paper yet underperform crude heuristics in production.
Digitalization has already displaced humans from the custody of data; the same drift now erodes their custody of information and knowledge. As volumes and refresh rates explode, the architecture of the systems—not the personal cleverness of their users—determines what the firm can or cannot decide well. Hence, understanding where information truly lives inside the company becomes strategic.
Since the 1970s, three software classes have structured that “where”: systems of records (run workflows, store transactions), systems of reports (describe what happened), and systems of intelligence (decide what should happen next). Most corporate disappointments come from conflating these roles—expecting ledgers to think, or analytics layers to serve as sources of truth.
This chapter proceeds in stages: we first disentangle data, information, and knowledge, anchoring “information” in Shannon’s theory. We then examine how these notions appear—and are routinely distorted—inside the three dominant enterprise‑software classes. With that frame, we confront the messy semantics of business data, the folklore of “bad data”, and the incentives that bend records away from the firm’s interest. Finally, we revisit the tacit, so‑called mundane knowledge that still sits in people’s heads and show how it, too, can be progressively codified. This sets the stage for the subsequent chapter on intelligence and for the full mechanization of decision‑making.
Data, information, and knowledge
Data, information, and knowledge are three words that presentation slides and vendor brochures happily interchange; in practice, confusing them wrecks architectures and policies. Treating raw records (data) as if they already resolved uncertainty (information), or treating reduced uncertainty as if it already carried a causal model that can steer action (knowledge), is not a minor semantic slip. It is the root cause of dashboards that never decide, “planning” modules that never plan, and post‑mortems that keep blaming “bad data” for failures that are, in fact, methodological.
Data are symbols captured by systems—bytes in tables, log lines, barcodes; by themselves, they say nothing about how uncertain we remain. Information is what those symbols buy us against uncertainty; it can be quantified and budgeted. Knowledge is the structure—explicit or tacit—that lets an intelligence turn information into decisions that systematically improve the firm’s returns. Collapsing these layers invites two symmetric errors. First, firms hoard ever more data and compute, convinced that volume substitutes for information; they end up with petabytes of low-grade exhaust that barely moves the needle on the uncertainty that matters. Second, they mistake information for knowledge and expect reports or KPIs to “drive” the business; people then compensate manually, injecting ad‑hoc rules and folklore that never get formalized, versioned, nor tested.
For a modern supply chain, the distinctions are operational. If you cannot say where information sits, you cannot say which system is authoritative; if you cannot say where knowledge sits, you cannot say who—or what—is accountable for the decisions. Once we look through this lens, “systems of records”, “systems of reports”, and “systems of intelligence” become essential architectural constraints: records hold (most of) the data, reports expose fragments of information, intelligence operationalizes knowledge to issue orders. Expecting any one of these layers to do the job of the others is wishful thinking—and a reliable way to burn both budgets and credibility.
What follows, therefore, does not rehash definitions for the sake of taxonomy. We start from the mechanical nature of data—mere symbols—and the engineering realities that govern their storage and retrieval. We then anchor “information” in Shannon’s quantitative treatment of uncertainty, showing why many seemingly “rich” datasets carry little informational weight for the decisions at hand. Finally, we turn to knowledge, the least tractable of the three, and tie it back to the company’s code, recipes, and people—where it actually lives, how it should be curated, and how it can be progressively mechanized. Only by keeping these three tiers distinct can a firm hope to build decision systems that are both informed and reliable.
Data storage
Data are recorded symbols. They can be numbers, characters, pixels, or any other discrete marks that a system can reliably store and retrieve. On their own, these symbols carry no meaning; only an agreed‑upon interpretation (a schema, a codebook, a protocol) turns them into something a firm can reason about. A barcode, a customer identifier, a purchase order line, or a temperature probe reading are all “data” in this strict, mechanical sense.
On computers, these symbols are ultimately encoded as binary digits and grouped into bytes (8 binary digits per byte)33. Modern storage devices, weighing just a few grams34, can hold over 1 TB, or $$10^{12}$$ bytes. A terabyte already lies beyond intuitive grasp. The 1768–1880 edition of Encyclopaedia Britannica spans 143 volumes, with its text content totaling 336 MB—about 3,000 times smaller than 1 TB. Even with high‑resolution images, the encyclopedia totals just 23.5 GB, or roughly 400 times smaller than 1 TB. With efficient representations, 1 TB can store over a decade of retail transactions for a large network.
From a supply chain perspective, large data storage capacities enable the collection of extensive electronic records. These records capture both the current state of the supply chain and the detailed sequence of events that led to it, including the flow of physical goods. They can also include market data, such as competitor pricing. All of these data are stored in the company’s systems of records, which will be discussed further.
The storage capacity of modern hardware is so vast that the numbers are hard to internalize—and this gap will only widen as technology progresses. Modern digital stacks routinely span 10 to 15 orders of magnitude: from nanosecond CPU cycles to day‑long batch jobs, from bytes on embedded devices to petabytes in data warehouses. By contrast, the span between a one‑person shop and the largest firms on Earth is “only” about seven orders of magnitude (1 to 10 million employees). The quantities are perfectly measurable; what is difficult is building an intuition for phenomena that differ by factors of a trillion or more.
As capacities rise, retrieving and processing what we store also become harder. Yet market pressure generally pushes the rest of the stack—bandwidth, RAM, CPU—forward enough to keep the extra storage economically usable. When storage outpaces its complements for too long, investment in ever‑denser media simply stalls until the rest of the hardware catches up, at which point further gains again make business sense.
A gram‑sized storage chip now holds more bits than a single mind—or a thousand of them—could ever keep in working memory. This sheer capacity is why, decades ago, business systems became the de facto custodians of almost all relevant supply-chain information. Consequently, a supply‑chain decision‑making process is truly informed only if it taps into the vast troves held by the company’s systems of records.
The vastness of modern data storage comes with two distinct pitfalls. Supply chain practitioners who are not well-versed in information technology often underestimate how much data can be stored and processed for supply chain purposes. As a result, vital data are left unrecorded, deleted too soon, or never collected. If these data were available, they could have been used to further refine decision-making processes. The two most common mistakes are failing to preserve all events affecting the business systems and discarding historical data after three to five years, even though the entire history could be preserved almost indefinitely35.
For example, while all inventory management systems include manual stock-level overrides to correct discrepancies between physical inventory and electronic records, some do not record those overrides. Yet the full trail of overrides is essential for tracing the root causes of inventory errors. Ideally, every event that ends up modifying the state of the business system should be recorded36.
Conversely, software engineers, who are well aware of computing hardware capabilities, often take advantage of this by favoring easy-to-implement solutions over resource-efficient ones. As a result, software often becomes inefficient in its use of hardware, especially in enterprise software compared to consumer applications. In consumer software, the buyer and user are typically the same person, and the app’s responsiveness is a key factor in its adoption37. Engineers therefore spend much of their effort squeezing every cycle out of the app.
By contrast, the buyer of enterprise software is seldom the user, and demos usually run on mock datasets far smaller than those in production, so real‑world performance gets little scrutiny. Moreover, despite inefficient use of computing resources, the total cost of ownership of the app usually dwarfs its hardware costs. As a result, when facing a performance issue, it is common to throw more hardware at the problem instead of improving the underlying efficiency of the software. More generally, most enterprise software vendors have only modest incentives to invest in making their technology resource-efficient.
Resource efficiency matters because the very same piece of information can be encoded in many ways, some far leaner than others. While a 1 GB ($$10^{9}$$ bytes) dataset may seem richer and more informative than a 1 KB ($$10^{3}$$ bytes) dataset, more bytes do not inherently mean more information. For example, the number 123 can be stored as a single byte or as a million bytes with a million leading zeroes.
Using a million bytes to represent a small number is an obvious misuse of computing resources, yet enterprise software often creates redundant copies of the same data in its databases. In relational databases, enterprise software typically inflates data storage by a factor of ten—and sometimes a hundred—due to redundant data representations. Various technical reasons justify these redundant data representations. Those reasons lie outside our scope here, but they raise an important question: how does the chosen representation relate to the information itself?
Information theory
Company records can be stored as data. Yet while data are conventionally measured in bytes, that unit does not indicate how much information they contain. The same record admits many representations, some more storage-efficient than others. The underlying information is invariant to its notation. For example, 1776 equals MDCCLXXVI in Roman numerals; the information is unchanged. Information theory studies what information is—and how it is measured, stored, and transmitted.
Claude Shannon founded modern information theory in the 1940s. It is a foundational auxiliary science of supply chain. This theory is grounded in probability and defines information as the capacity to resolve uncertainty. To illustrate the link between uncertainty and information, consider two stores, each stocking 100 products. We focus on stock levels—100 numbers in total—and compare the information involved in managing two very different stores.
First, a luxury watch store. For each watch in the assortment, at most one unit is kept. When a watch sells, a replacement is promptly shipped to keep the assortment complete. Second, a suburban depot for bulk construction materials. Replenishments are infrequent and involve large batches, typically a full truckload. In the watch store, a diligent manager likely knows the exact inventory at all times. Typically, the stock level is 1 for all watches; for those just sold, it temporarily drops to 0 until replenishment. Because most watches are at 1 and only a few briefly at 0, the manager can easily recall levels. In contrast, stock levels at the suburban depot range from 0 to 10,000 units. A manager might recall which products are out of stock or recently replenished, but even a diligent one cannot memorize the exact levels across all 100 products. Both stores track 100 items, yet the watch shop carries far less information than the depot.
Information theory makes this contrast precise. The notion of entropy quantifies information; it is measured in bits (sometimes “Shannons”). Returning to the example suffices to illustrate entropy. Assume the watch store has a 95% service level: each watch is out of stock with probability 5%, otherwise at 1. Assume further that the suburban store’s stock levels are uniformly distributed between 0 and 9999. Although a uniform distribution is unrealistic, it suffices for this illustration.
The entropy will be approximately 29 bits for the luxury watch store and 1328 bits for the suburban depot (the detailed calculation appears in the annex). The suburban depot’s entropy is nearly $$50\times$$ that of the watch store, which is why even a diligent manager cannot mentally track its inventory.
From a supply chain perspective, information theory clarifies what complex means in a supply-chain system. A supply chain resists succinct description: millions of bits are needed to represent even an average one, and billions for a large one. This vast requirement is not an artifact of representation; it is a fundamental property of supply chain. The information required to resolve the state of a supply chain far exceeds what any group of people could memorize, let alone track its transitions over years.
Thus, generating informed supply-chain decisions daily requires millions of bits as input to reflect state, and millions more as output to represent the decisions. Hence software is fundamental to modern supply chains: without it, operating them would be impossible. Before business systems spread in the early 1970s, supply chains were far less complex than today. This surge in complexity stems directly from advances in computing hardware. Companies now manage more SKUs, suppliers, sites, price schemes, and channels than ever, leveraging modern computing hardware to handle this increased complexity.
Although supply chains have grown more complex, many sources of accidental complexity have been tamed. For physical goods, standardization, strict tolerances, and formal quality control became increasingly common in the first half of the 20^th^ century. The digital turn of the 1970s pushed standardization even further. The modern concept of SKU (stock-keeping unit) assumes that all units in stock are physically identical.38 The content of an SKU can be summarized by a single number: the stock on hand.
Modern readers may find it hard to believe that even mass-produced items in the 19^th^ century varied markedly from unit to unit. In such supply chains, an inventory management system—had it existed—would have needed to track details for every unit to reflect inventory accurately. In practice the problem was handled differently: salesmen discounted sub-par units on the spot.
Information theory provides fundamental insights for supply chain, especially when judging whether a data source can sharpen decisions.
For example, this perspective clarifies why fine-grained manual overrides—such as adjusting forecasts or stocking policies—are futile. If demand planners rely on historical data to make those corrections, they add no bits of information to the system. They merely palliate defects in the existing numerical recipes—built on that very same data. Thus, for such a manual process to make the system more informed, it must introduce “fresh” bits of information. Suppose a number based on intuition or prior knowledge carries fewer than 10 bits of information. Even if an employee supplies 100 meaningful numbers daily, this totals only 1000 bits—assuming, unrealistically, complete independence. Adding 1000 bits to a system that processes millions is unlikely to make it materially more informed. While an employee may occasionally supply information with outsized impact, this raises questions about the source of such insights and the odds of repeating them consistently.
More generally, the informational lens helps assess how far a data source can inform supply-chain decisions. As the earlier example showed, processing data adds no information—it can only uncover what was already there in the original data. It also clarifies that data volume may not correlate with information content.
Yet turning information into informed decisions still requires one more ingredient: knowledge.
Knowledge
Knowledge is the theoretical and practical understanding of an object. It encompasses the insights, rules, and principles that shape a mind’s causal model of the object39. This causal model lets a mind deduce why an object is as it is and where it is headed. Knowledge turns observations into actions that steer the object toward a better state.
For supply-chain purposes, we draw on only a tiny fraction of mankind’s knowledge. Our concern is solely with supply chains; the scope of knowledge is limited to this purpose. Knowledge counts as supply‑chain knowledge only when it improves the company’s decisions. Supply-chain knowledge, in the broadest sense, spans a wide range of subjects. However, as the earlier Chapter Epistemology explained, we must distinguish between the core supply-chain domain and its auxiliary sciences, as their ways of organizing and fostering knowledge differ markedly.
Unlike information—fully characterizable mathematically and an unintelligent mechanical affair whose storage medium is irrelevant—knowledge is elusive and inseparable from the mind that produces it.
Give two students the same physics manual: one may extract insights far beyond its examples, while the other may never connect examples to principles. What a reader gains is subjective: a book’s knowledge remains latent until someone reads it. Until then, it is merely a data storage device. The book’s content remains the same regardless of paper quality—or even of being printed at all; the electronic copy is identical in content.
Because knowledge and general intelligence are intertwined, fully understanding one requires fully understanding the other. Defining one without the other is inevitably incomplete. General intelligence remains a subject of active research. Progress in this area has accelerated over the past seven decades, yielding remarkable results. Yet we have also realized how much we underestimated our ignorance on this subject. We will revisit this topic in Chapter Intelligence.
Equally important is the locus of knowledge—where it actually lives inside the company—a crucial question when applying knowledge for supply chain purposes. Two common but simplistic views miss the mark. The first and most common misconception is that knowledge resides exclusively with experts and their supporting materials—papers, books, and courses. Academia often feeds this view; many professors see themselves as the pinnacle of expertise. The second is the notion that knowledge is spread evenly across the company. This belief often stems from a misguided desire to involve all employees in the company’s supply chain efforts. To clarify the locus of knowledge, we must further refine what we mean by knowledge.
In his landmark 1945 essay40, The Use of Knowledge in Society, Friedrich Hayek separates ‘special’ from ‘mundane’ knowledge. Special knowledge covers textbooks, formulas, source code, policies, and regulations. The fundamental error, he argues, is to believe that special knowledge is the only type.
Hayek adds a second, crucial kind—mundane knowledge. Mundane knowledge covers the trivia bound by the circumstances of a given time and place. A truck driver, for instance, knows the healthy hum of an engine and books repairs before trouble starts. This knowledge is sharpened when the driver regularly operates the same truck, developing an understanding of its unique quirks over time. This knowledge is informal and undocumented, yet often essential for getting things done reliably.
Thus, supply chain mixes special knowledge, held by a small body, and mundane knowledge, held by a much larger one. In either case, the holders of that knowledge may be employees—or not. For example, bits of special knowledge can be delegated to an enterprise software vendor, and bits of mundane knowledge to a third-party repair company.
For a century, firms have worked to replace mundane know‑how with codified rules. Mundane knowledge still matters—what you don’t know can hurt you—but its informality creates friction for large firms. Large companies aim to make their processes as replicable as possible to reduce accidental variance in quality, cost, or delays in their physical-goods flow. From this perspective, mundane knowledge hinders the standardization of processes.
The rise of enterprise software over the last five decades has amplified this trend: software embodies special knowledge and accelerates the shift. Software mechanizes the application of knowledge, yielding significant productivity gains when successfully deployed. However, software is expensive to set up and even more costly to maintain, especially in complex enterprise settings. Codifying a process yields two gains: easier replication and tighter software control.
In a modern supply chain, the source code that implements the firm’s numerical recipes—together with its supporting artifacts (tests, notebooks, configuration, documentation)—is the company’s most valuable special knowledge asset. Unlike slideware or tribal lore, this asset is actionable: it turns information into decisions without waiting for a meeting. Crucially, this codebase is not a static library of formulas; it is a living corpus that must be curated, versioned, and continuously challenged against profit-and-loss feedback.
Because the codebase is information, it can itself be analyzed by software. Static analysis, type systems, linters, profilers, formal tests, and simulation harnesses make the knowledge both inspectable and improvable at machine speed. Conversely, because this knowledge is inseparable from the minds that evolve it, the engineers who maintain and extend the codebase are part of the asset: without stewardship, rot sets in, semantics drift, and the firm quietly slides back toward folklore.
This observation closes the loop. Data are symbols. Information is their power to shrink uncertainty. Knowledge is the causal structure—captured increasingly as code—that systematically converts that reduced uncertainty into superior allocations of scarce resources. For supply chain, “owning the knowledge” means owning and operating the code that embodies it.
Classes of enterprise software
Most of the information a modern supply chain needs resides in enterprise software. Yet these applications play distinct epistemic roles. The previous pages separated data, information, and knowledge; this section mirrors that tripartition in software. What the firm records (data), reports (information), and decides (operationalized knowledge) live in three distinct classes of systems. Confusing them—asking ledgers to think or dashboards to be sources of truth—is among the costliest mistakes companies make.
Since the 1970s, three classes have dominated enterprise stacks: systems of records, systems of reports, and systems of intelligence. Dozens of specialized tools exist (e.g., CAD for engineering, virtualization for IT), but in budget and supply-chain relevance they are dwarfed by these three. What follows names each class precisely, explains why their responsibilities must remain disjoint, and shows how mixing them corrupts performance and accountability.
Its primary virtues are reliability, traceability, and consistency. “Intelligence” is neither required nor desired. A system of records digitizes and streamlines tasks otherwise done with pen and paper. Almost all transactional systems—ERP, CRM, WMS, PLM, MRP—belong to this class. Architecturally, they are CRUD-first (create, read, update, delete) and backed by a relational database. Concurrency control, auditability, and strong consistency dominate the design. Heavy analytics and optimization do not belong here: they starve shared resources, degrade the user experience, and muddy the accountability ledgers must preserve.
It adds no information; it merely reshapes what exists. Systems of reports automate what clerks and analysts once did by hand: totals, averages, pivots, dashboards. They sit on top of systems of records and are intrinsically descriptive. For modern supply-chain practice, their usefulness is limited: they remain helpful for compliance, forensics, and management oversight, but not as inputs to optimization engines. Running decision-making on a reporting layer introduces information loss, semantic drift, and extra bias for no economic gain.
It does so through a narrow API and exposes logs precise enough to be challenged against profit-and-loss feedback. It owns almost no master data and very little user interface. A system of intelligence replaces the manual, spreadsheet‑mediated steps that typically sit between reports and execution. It consumes authoritative data, computes options, arbitrates them economically, and commits orders, allocations, prices, or schedules to the ledger. Its engineering culture, runtime profile, and accountability model differ fundamentally from those of systems of records and reports. This class will be examined in greater depth in Chapter Decisions.
Inside a firm, systems of records come first, reports second, intelligence last. Systems of reports and systems of intelligence are intended to be overlaid on systems of records—possibly aided by data lakes or warehouses. Companies often err by blending the classes, expecting a single product to deliver value beyond its natural boundaries; those attempts invariably fail. Systems of records make terrible systems of intelligence, and vice‑versa.
The second mistake is attempting to run a system of intelligence on top of a system of reports. A system of reports is a secondary source of information. The original data are invariably altered to make it more amenable to reporting. These transformations not only cause information loss but also introduce complications and bias. While those transformations are justified—within the context of the system of reports—the data no longer accurately reflect what actually happened.
For example, weekly sales reported by business intelligence may be net of returns. For items often sent back, those figures can understate the inventory required to maintain a given service level.
The idea of running a system of intelligence on top of a system of reports is tempting because, to human eyes, the data as presented by the system of reports looks clearer and better organized than its original counterpart in the system of records. Thus, it feels as if this would help the system of intelligence. Yet this is false. The information perspective explains why: the system of reports not only contains less information but also adds its own layer of complications.
A clear grasp of these three software classes is essential for modern supply-chain practice to take root in a company.
The place of software
Supply chain is now software-bounded: what the firm can observe, remember, and ultimately decide is constrained less by forklifts and bays than by schemas, APIs, and runtimes. Misplacing software in the overall architecture is not a pedantic IT concern; it is an economic error that quietly taxes every decision the company takes. If the wrong system is asked to think, or the right one is starved of the data it needs, optionality shrinks, variability bites harder, and working capital gets trapped for reasons that never show on a P&L.
The previous pages separated data, information, and knowledge, and mapped them to three software classes: systems of records, systems of reports, and systems of intelligence. This section clarifies why that separation must be enforced in practice, not merely acknowledged on slides. Before examining how most writers get this wrong, we must make explicit what is at stake: authority over the facts (records), over their narration (reports), and over the levers that actually move cash and atoms (intelligence). Confusing these roles scrambles accountability, bloats costs through resource contention, and ensures that “bad data” will be blamed for what is, in truth, a design failure.
With this frame in mind, we can look at how supply-chain discourse routinely miscasts software—either as an irrelevant detail that magically provides whatever numbers a formula demands, or as a glossy org chart of interchangeable modules that supposedly add up to “best practice”.
Supply‑chain writers typically fall into two camps. Those with an academic bent treat enterprise software as an abstraction, assuming that figures such as sales or stock levels are simply “there”. Where the numbers come from is seldom discussed; obtaining them is treated as casually as a contractor choosing a shovel brand. To widen their scope, they offer a menu of models—some that rely on, others that sidestep, inputs such as stock-out records—leaving firms to pick whichever fits the data they actually hold.
Writers from consulting firms lean on fashionable buzzwords, unfurling ornate diagrams and maturity matrices that treat software modules as puzzle pieces. They prescribe phased “adoption journeys”, but the stages seldom get more than slide-deck gloss. Executive sound‑bites abound; hard technical substance is scarce.
Both viewpoints fall short of a rigorous supply‑chain treatment. Treating “sales data” or “stock data” as self-evident betrays ignorance of enterprise software realities; describing tools without their underlying mechanics or markets is no better. Too often the result is little more than a sales pitch in scholarly dress.
Software is an indispensable auxiliary science of supply chain, and its mastery cannot be shunted to specialists—no more than a modern physician can outsource chemistry. A firm must understand its own software terrain to run its supply chain well.
Systems of records
Systems of records are a company’s ledgers: they embody operational workflows, capture every transaction and state change, and, by construction, become the authoritative memory of goods in motion. In modern supply chains, they are the only scalable custodian of information: what the firm can observe and remember sits there—not in people’s heads.
This section treats systems of records as they are: indispensable, yet epistemically and technically bounded. They are engineered for reliability, traceability, and concurrency control. They are not built to report, let alone decide. Confusing them with systems of reports or systems of intelligence is among the most common—and most expensive—corporate mistakes.
We proceed in five steps: first, we recall the historical path that produced today’s ERPs, CRMs, WMSs, and their cousins. Second, we dissect their CRUD design and the ensuing commoditization. Third, we examine the technological trajectory that makes every system of records swell over time, creating functional overlaps and synchronization nightmares. Fourth, we state the hard functional limits that preclude pushing heavy analytics or optimization inside the ledger. Fifth, we close with the vendor’s conundrum—the structural incentive to market ledgers as “intelligent” platforms despite their intrinsic bounds.
With this frame in mind, the rest of the section can be read as a field guide: how to exploit systems of records ruthlessly for what they do well, and how to keep them out of the decision seat they were never designed to occupy.
The origin of enterprise software
Systems of records appeared in the late 1970s and spread rapidly through the 1980s and 1990s. Their role was to convert paper-heavy clerical work—especially accounting and stock keeping—into electronic workflows. By capturing every transaction and stock movement, they eliminated whole classes of clerical error and, once barcode scanners arrived, lifted productivity. By the late 1990s, most large firms had deployed several such systems to automate their routine back-office tasks.
Falling hardware costs underpinned this rise. During 1970–2000, Moore’s Law held: transistor counts doubled roughly every two years while prices barely moved. Memory, storage, bandwidth, and I/O followed similar curves, so by the late 1990s storing transactions digitally was orders of magnitude cheaper than paper alternatives.
Amid the flood of record‑keeping products, two labels came to dominate: ERP (enterprise resource planning) and CRM (customer relationship management). Strip away the marketing—both are the same animal: a CRUD ledger wrapped in workflows. The only real difference is the entry point—ERP is organized around the product (and its bills of materials, routings, inventories), CRM around the customer (and the sales pipeline). Once a vendor secures a foothold, both families grow the same way: bolt on adjacent entities, bundle ever more modules, and acquire niche rivals. By the late 1990s, most ERPs and CRMs had swollen into sprawling monoliths, setting the stage for the technical problems discussed next.
One root cause was technical: packet-switched networking was conceived in the 1960s, but true distributed business systems stayed niche until the late 1990s. Firms therefore ran centralized mainframes, and cross‑vendor integration was either ruinous or impossible. Expanding the resident system was cheaper than wiring in a new one.
But building and running ever‑larger code bases is expensive. During the 2000s commodity integration tools—open‑source ETLs and web APIs—made it cheaper to stitch systems together, so specialized record‑keeping apps resurfaced and clawed work back from the monoliths.
Yet the old monoliths never shrank, and the new niche tools soon ballooned too: every system of records keeps growing along the same lines. Hence the alphabet soup of WMS, OMS, PIM, PLM, SRM, and countless cousins—different starting points; much the same architecture and growth path.
In supply-chain work, the ERP monolith is typically the prime data source because it holds the purchase and sales ledgers. The label Enterprise Resource Planning is a marketing artifact: an ERP does not plan in the economic sense of arbitrating options under uncertainty; it executes and records resource workflows (orders, inventories, invoices) and, at best, sequences clerical steps (approvals, calendarized MRP explosions, status flips) according to user‑entered parameters.41 Calling this “planning” conflates bookkeeping with decision-making and fuels the recurring myth that a system of records can double as a system of intelligence.
The CRUD design
Nearly every current record system follows the CRUD (create, read, update, delete) pattern born with relational databases in the late 1970s42. Front‑ends have evolved from green screens to glossy web apps, yet back‑end principles have scarcely budged. Tooling, by contrast, has advanced so far that, since the mid-2000s, systems of records have been little more than commodities.
In practice “the database” means a relational (SQL) engine: a set of
tables, each with typed columns (numbers, text, dates, …) and rows of
data. Other designs exist but matter far less in enterprise software.
Relational databases link tables through primary‑ and foreign‑key pairs.
An Orders
row, for instance, can carry a ProductID
that points to
the matching ProductID
in Products
; the engine forbids an order that
references a product that does not exist.
Vendor dialects differ, but the SQL‑92 (published in 1992) core works everywhere, and most CRUD apps rarely need more than that despite later, fancier extensions. The fine print of SQL is beyond the present scope; however, a basic level of SQL should be considered part of the foundational knowledge needed by a modern supply chain practitioner.
CRUD design begins by defining entities—groups of tables mirroring the
business concepts the app manages. Those concepts can be products,
clients, inventories, suppliers, contracts, employees, payments, etc.
Moreover, most business concepts do not fit within a single table. For
example, a client order may require one table Orders
with one
transaction per line, and another table OrderDetails
listing all items
included in each order.
Each entity gets screens through which users create, read, update, or delete its records, following the database’s underlying paradigm. Those screens are frequently accompanied by bits of business logic to automate database changes—mechanical consequences of the user’s entries—and typically follow a workflow of some sort.
A CRUD app’s complexity almost always creeps upward: new entities appear; existing ones expand. New screens proliferate; old screens accrete options. Workflows grow more expressive by adding options that fine-tune the automation that unfolds after data entry. In theory, complexity could fall; in practice, ambient incentives prevent it. We return to this point next.
CRUD apps are, as the name suggests, crude. Putting aside the foundational layer itself (database included), there is nothing remotely “smart” in either the design or the execution of a CRUD app. The software principles supporting the rules and behaviors implemented in a CRUD are fundamentally accessible to anyone who has completed middle school. It boils down to arithmetic operations interleaved with IF-THEN statements, and little more. Naturally, the business itself may be intricate, with rules and behaviors well beyond what a middle-school student can comprehend. In such instances, the CRUD paradigm offers only a laborious decomposition into elementary processing steps.
CRUD’s simplicity made it the default business‑app pattern for decades, and since the mid‑2010s open‑source stacks have turned it into a commodity. Those tools make software engineers far more productive when developing CRUD apps than when developing almost any other kind of software. Notably, CRUD apps are the poster child in most programming-environment demonstrations. Moreover, while hiring software engineers remains notoriously challenging, CRUD development is largely accessible to junior developers (and to mediocre seniors) and represents, by far, the easiest kind of software to build.
Many firms still overpay for CRUD record systems, underestimating how thoroughly open‑source tooling has commoditized them. Vendors exploit that lag to keep prices high. No matter how useful and mission-critical a system of records might be, this sort of product can now be developed both quickly and cheaply. Yet, out of entrenched habits, many companies still pay software vendors annual fees beyond what it would cost to redevelop such systems in-house. Naturally, vendors benefit from this confusion and have become masters at delaying the inevitable: lowering prices on the ledgers they sell. This situation also follows from the technological trajectory of systems of records.
Technological trajectory
Systems of records, unlike the spreadsheets they replaced, are rigid and less expressive. Even an extremely capable system of records today covers only a fraction of what a loose collection of spreadsheets supports. With loose spreadsheets, employees can always deviate from formalized processes—so long as their colleagues agree. Steps or requirements can be skipped; exceptions can be made to any rule; or, conversely, an arbitrary judgment call can reject a request that “blind” policies would deem acceptable. Naturally, this freedom costs reliability and productivity, as spreadsheets resist substantial automation.
Vendors typically launch a system of records with a small set of entities plus the screens and workflows needed to handle them. Technological resources at an emerging vendor are limited, and there is no point in further development if the product can be sold as is despite its limitations. Furthermore, developing a business app without detailed feedback is futile before customers have embedded it in their processes. Thus, while the initial core may rest on an a priori view of the market, extensions are almost invariably built in close collaboration with a few customers.
Business markets lack the depth of consumer markets. As a result, once a product is sold to a business client, the vendor naturally leverages that position to sell anew to the same client—upselling43. Selling extra modules to an existing customer is far cheaper than winning a new one, since the latter entails clearing all corporate hurdles of a new line of spending—such as going through a tender against rivals.
Systems of records always offer another entity to add—or enrich—so the vendor can justify the next sale. No matter the product’s feature coverage, there are always more business concepts and workflow variations to add. Nothing in the technology of systems of records imposes a consistency imperative that would restrain growth. For example, an inventory management system can be expanded to manage customers; conversely, a customer-relationship-management system can be expanded to manage inventory. In fact, most large ERP vendors feature CRM capabilities; conversely, most large CRM vendors feature ERP capabilities. It boils down to adding more tables, more screens, and more business logic, ad infinitum. Setting aside per-client bespoke customization, the system of records becomes the sum of the quirks of its entire customer base.
Every commercially‑maintained system of records keeps growing until the vendor goes bankrupt or retires the product line entirely. As engineering cost is a superlinear function of the number of entities—i.e., implementing twice as many entities more than doubles the costs—the product’s growth rate (in features and capabilities) tends to slow over time, even if the vendor keeps increasing its engineering headcount as the company expands. Yet, despite those superlinear costs hindering an otherwise unlimited complexity growth, many systems of records end up as feature mastodons after two decades of continuous development. Numerous popular systems of records, including those for smaller businesses, boast over 10,000 tables, most with dozens of fields.
The path of relentless complexity growth is effectively irresistible for vendors selling systems of records. There is simply no upside in ditching infrequently used entities: it antagonizes the few client companies that use them—however small that number may be—while leaving all other clients indifferent. Unfortunately, this path leads to a predictable bastardization of the product. Indeed, while collaborating with customers is an essential ingredient of business software development, some customers are misguided in their requests. Others reflect narrow niches that did not warrant bespoke capabilities. Still others impose terminologies or concepts inconsistent with those adopted by the vendor. As a result, over time, every actively developed system of records becomes a Lovecraftian construct endangering the sanity of those who contemplate it for too long.
In A Deepness in the Sky (1999), American science-fiction writer Vernor Vinge offers an uncannily accurate picture of long-running software initiatives:
There were programs here that had been written five thousand years ago, before Humankind ever left Earth. The wonder of it—the horror of it, Sura said—was that unlike the useless wrecks of Canberra’s past, these programs still worked! And via a million circuitous threads of inheritance, many of the oldest programs still ran in the bowels of the Qeng Ho system. Take the Traders’ method of timekeeping. The frame corrections were incredibly complex—and down at the very bottom of it was a little program that ran a counter. Second by second, the Qeng Ho counted from the instant that a human had first set foot on Old Earth’s moon. But if you looked at it still more closely. . .the starting instant was actually some hundred million seconds later, the 0-second of one of Humankind’s first computer operating systems.
So behind all the top-level interfaces was layer under layer of support. Some of that software had been designed for wildly different situations. Every so often, the inconsistencies caused fatal accidents. Despite the romance of spaceflight, the most common accidents were simply caused by ancient, misused programs finally getting their revenge.
“We should rewrite it all,” said Pham.
“It’s been tried,” corrected Bret, just back from the freezers. “But even the top levels of fleet system code are enormous. You and a thousand of your friends would have to work for a century or so to reproduce it.” Trinli grinned evilly. “And guess what—even if you did, by the time you finished, you’d have your own set of inconsistencies. And you still wouldn’t be consistent with all the applications that might be needed now and then.”
Yet this spiral of accretion is not an iron law for the buyer. Several large organizations have kept their ledgers lean by pushing ancillary workflows into detachable sidecars—small services or plug-ins that interface with the core through an API fence. Because the plug‑ins are owned, versioned, and—crucially—disposed of by the client at will, cruft accumulates at the edge where it can be pruned without touching the authoritative data store.
A more radical lever is phased retirement of functions. Contrary to the folk wisdom that “nobody rewrites an ERP”, many firms do—just not in a single heroic blast. The proven pattern is the strangler: place a thin façade in front of the monolith, route a single, self-contained capability (pricing engine, returns authorization, tax module …) to a greenfield service, freeze the corresponding tables, then physically drop them after a sunset period. Repeat every quarter and the impossible big‑bang becomes a sequence of manageable surgeries.
These tactics do not fix the vendor’s incentive to sell yet more expansion packs; they merely restore agency on the client side. Nothing forces a company to install the next glossy add‑on when the carrying cost of complexity exceeds the cost of a lean replacement. In practice, the lowest long‑term cost of ownership comes from treating the central ledger as sacred ground and everything else as disposable code. Complexity growth is the default, not the destiny—provided the organization cultivates the discipline to keep pruning its own garden.
Functional overlaps
Master data44 is the first casualty of functional overlaps. As soon
as two systems of records coexist, no 1-to-1 mapping of entities can be
assumed. Attributes that look identical (“product weight”, “order
date”, “supplier code”) routinely diverge in meaning, unit, timing, and
governance. What the ERP calls product_family
may only approximate
what the PIM stores as category
; lead_time
in the MRP may be a
procurement promise while the WMS records an empirical average. The
so‑called “golden record” is a managerial fiction: in reality, each
system is master for some attributes, some of the time, for some
use‑cases.
This fragmentation is not a mere clerical nuisance; it is the primary driver of semantic drift (see Section Semantics of the data). When overlapping ledgers do not share isomorphic schemas, information about a single business fact gets smeared across tables, systems, and teams, with no canonical reconciliation rule. Master‑data‑management (MDM) programs attempt to paper over this reality with central hubs and survivorship rules, but—absent a clear, per‑attribute authority and an auditable decision system—they merely add another overlapped system and another layer of synchronization to maintain.
Functional overlaps between third-party systems of records are the direct consequence of their technological trajectory, and they hit master data the hardest. Indeed, as software vendors keep pushing their respective functional envelopes, collisions are inevitable; only their timings remain uncertain. Moreover, larger vendors are especially prone, having grown by continually widening their functional coverage. This problem is specific to third-party systems; even modest internal coordination usually prevents such overlaps from arising in in-house developments.
In practice, overlapping systems require data-synchronization processes. Master‑data synchronization is particularly pernicious: schemas are rarely isomorphic, attributes carry different semantics, and per‑attribute authority (which system is master for what, and when) is seldom spelled out. Such synchronizations are invariably ad hoc software projects. Indeed, the enterprise‑software market is too fragmented to justify developing plug‑and‑play synchronization tools.
A two-way synchronization between systems of records—where edits can be made on both systems and each edit in one must be propagated to the other—typically requires at least an order of magnitude more effort than one-way integrations, where one system is deemed “master” and the other “slave”. A two-way data synchronization must address conflicting edits and their resolution, which is far trickier than simply copying and overwriting all edits from the master into the slave. For master data, where multiple departments continuously edit different subsets of attributes, this conflict surface is the rule, not the exception. Despite the existence of exceedingly clever algorithms dedicated to the case, N-way data synchronization, with N greater than 2, should generally be regarded as an abomination.
Even when IT carefully identifies which system is master for each area, most large systems end up both master and slave—depending on the area (e.g., the ERP may be authoritative for price and the PIM for descriptive content on the same SKU). For example, the CRM may be master for “customers” yet slave to the ERP for “suppliers”.
These overlaps pose a massive, ongoing hurdle that companies consistently underestimate. In particular, when it comes to picking the next software vendor, companies often prize impressive feature lists, as illustrated by the typical RFQ (Request For Quotation) process that frequently includes hundreds of requirements. As a result, companies tend to pick large vendors with needlessly complex products, following the common-sense intuition qui peut le plus, peut le moins (he who can do more can do less). Yet this misguided intuition invariably leads companies to pick vendors with vast functional overlaps. Those overlaps prove extremely costly.
These overlaps are not just an IT tax; they directly corrupt meaning. We return to this point in Section Semantics of the data, where we show how elusive, row‑dependent semantics quietly derail optimization efforts.
First, for the IT department, resolving synchronization issues is sharply superlinear—the effort grows far faster than the number of entities. Soon enough, the Tokyo subway map looks simpler than the synchronization graph that IT has to maintain. It is now common to see IT divisions go effectively bankrupt: they can no longer keep up with maintaining the synchronization graph amid the never-ending stream of updates to their third-party systems of records.
Second, for employees entering data, these overlaps create an ongoing source of confusion, as the same entities can be viewed and edited in more than one place. Software vendors have no incentive to facilitate the “enslavement” of parts of their products to rivals in order to mitigate this confusion. From their perspective, even an unused feature keeps a foot in the door for the next upsell. Furthermore, the idea that the company ought to pay for having features removed from a system simply goes against the entire paradigm governing the enterprise software market.
Third, all attempts at exploiting the data, either as part of a system of reports or a system of intelligence, confront a hall of mirrors: the same information appears in multiple places. However, secondary data, as generated by synchronization processes, should not be used for those purposes. Secondary data might be outdated, degraded, or subtly altered, wreaking havoc in downstream processes. Thus, for every piece of information, one must ensure he reads the canonical master record.
While the history of enterprise software is short, in-house development of systems of records—despite their flaws—is the only viable way to avoid the relentless costs induced by functional overlaps between third-party systems of records. Fortunately, as noted above, developing such systems of records has, for all practical purposes, been commoditized.
Functional limits
While systems of records deliver value to the firms that run them, what they can deliver is strictly limited. Systems of records are, at best, improved ledgers. In particular, they are not effective systems of reports nor systems of intelligence; those aims are technologically antagonistic to the basic functions they must serve.
Much of a system of records’ value derives from its database—its transactional core. The system’s design is fundamentally centralized—precisely what the business wants. For example, if the system holds the stock level of a given SKU, every user or agent querying it must see the same value. Furthermore, if someone claims 1 unit, that unit must not be granted to a competing claim. Maintaining electronic-stock integrity requires the system to coordinate competing concurrent claims on the same datum—a unit in stock.
More generally, this class of issues is resource contention: conflict over access to a shared resource such as RAM, disk, or other hardware. Resource contention is notoriously hard, as it often boils down to reducing latency. Yet latency progress largely stalled over a decade ago—over two for many specific challenges. To a large extent, latency is now bounded by the speed of light itself. For example, data packets over the Internet propagate at roughly two‑thirds of the speed of light in vacuum. This leaves little room for further improvement45. Another example: in about 0.3 nanoseconds—roughly a single CPU clock cycle—light can only travel back and forth about 5 cm.
Techniques exist to mitigate resource contention on computationally heavy processes—horizontal partitioning (sharding), read/write splitting, asynchronous queues, CQRS46/event sourcing, distributed locks, saga/orchestration patterns, and more. Each, however, trades the very virtues a system of records must maximize—simplicity, traceability, strong consistency—for operational and semantic complexity: eventual consistency to explain downstream, duplicated read/write paths, elaborate retry/compensation logic, brittle distributed transactions, heavier observability burdens, slower schema evolution, and harder audits. In short, a ledger can be made to scale with such patterns, but the price is ambiguity and fragility precisely where the firm needs an authoritative, crystal-clear truth. Consequently, anything computationally “fat” should be expelled from the ledger and delegated to a system of reports or of intelligence.
Resource contention largely explains why business apps are not fundamentally snappier today than in the early 2000s. Editing an entity still incurs a user-noticeable delay—tens of milliseconds. Lags of a second or more for seemingly basic operations remain common. Such lags seem surprising given that, at equal cost, hardware is now orders of magnitude more capable than two decades ago. Yet the stagnation reflects flat latency progress; resource contention does not improve either.
Inside the system of records, because resources are shared, every operation must be as light as possible. Any heavy operation risks starving the system, degrading the quality of service for other concurrent operations. Hence decision-making processes must be scrupulously avoided within a system of records. Decisions require weighing numerous options and entail orders-of-magnitude more compute than basic arithmetic or logical rules.
Consider a procurement optimizer that needs 10 seconds of CPU to compute the order quantity of a single SKU. In a system of intelligence, run offline or on a dedicated cluster, that cost is trivial—and even 100 seconds to squeeze a few extra basis points may be rational. Inside a system of records, however, the same computation is catastrophic: with tens of thousands of SKUs per day, 10 seconds per SKU becomes dozens of CPU-hours that contend with transactional workloads, stall screens, and degrade user experience. Even at 1 second per SKU, it would still create unpredictable latencies where the ledger must guarantee low, stable response times. Heavy optimization therefore belongs outside the ledger: the system of records should persist compact outcomes (e.g., purchase-order lines), not perform the combinatorial search that produced them.
In theory, alternatives to CRUD exist to interleave heavyweight with lightweight operations. In practice, however, those alternatives undermine the very reasons CRUD is attractive. The simpler, more practical approach is to segregate all the “fat” operations away from the system of records. Descriptive statistics belong in systems of reports; decision-making belongs in systems of intelligence. Keep the three classes strictly decoupled, and resource contention is largely mitigated.
Beyond low-level contention lies a weightier obstacle: the engineering cultures of the three software classes are fundamentally incompatible. From afar every class appears similar—each involves writing code—but moving between them is as different as house painting is from fine art. Both trades use paint; they share little else. Detailing those nuances is outside our scope; suffice it to say that for five decades vendors have strayed beyond their home class—most often with dismal results, especially when systems-of-records specialists expand into systems of intelligence. Yet vendors keep trying, oblivious to the predictable failures that follow.
Vendor’s conundrum
The value-add of a system of records is fundamentally limited. Once the system has ample coverage and proves reliable, no further benefits accrue. Software vendors selling systems of records have always been painfully aware of this limitation. Hence, since the late 1970s, nearly every successful vendor of systems of records has tried to venture into systems of intelligence.
To understand why, let’s consider the case of an accountant. There is little tangible benefit in a great accountant over a merely competent one. Hence companies do not hire “superstar” accountants at hundred-fold salaries; accounting delivers only bounded value. Marketing, by contrast, can add value without limit. A marketing genius can revitalize a brand in seemingly impossible ways, creating fortunes for shareholders. As a result, companies routinely seek “superstar” creative minds and pay them lavishly.
A system of records contributes like an accountant. Once the job is done properly and diligently, nothing more remains to expect. By contrast, the value-add of a system of intelligence—the decision-making engine—is akin to that of marketing talent. For practical purposes, there is no upper bound on the quality of a business decision.
Consequently, willingness to pay for newer, better systems of records is limited. This fundamental insight isn’t novel and was already perceived as such by software vendors in the 1970s. Until the early 1990s, successful vendors had little engineering capacity beyond adding features and keeping up with steady hardware progress. This changed in the 1990s and led many to rebrand as “ERP” vendors, the novel “P” for planning signaling an intent to be systems of intelligence. Similar moves occurred across other variants of systems of records. Yet, as noted, those attempts have largely failed—and still fail—sometimes spectacularly.
As a side effect, vendor marketing often confuses executives. As the distinction between systems of records, systems of reports, and systems of intelligence remains relatively unknown to the general public, vendors invariably tout the promise of better decisions. In short, every enterprise software product is pitched as a system of intelligence. For instance, a system that merely logs stock movements while clerks handle replenishment should not promise fewer overstocks or stockouts—yet such claims are ubiquitous.
From the buyer’s angle, three simple heuristics cut through the confusion.
First, a bona fide system of intelligence ships with almost no data-entry screens. Intelligence consumes authoritative records; it does not solicit daily re-keying or “confirmation” by users. Each additional form is a tell-tale sign of a ledger in disguise, and it muddies accountability: when outcomes are poor, the vendor can always claim that “the user typed the wrong number”. A genuine system of intelligence owns its decisions end-to-end: its inputs are the firm’s records; its outputs—orders, allocations, prices, schedules—are auditable against profit and loss. No daily clerical escape hatch remains for either party.
Second, the engine must not pretend to be a standalone solution. Beyond its own code, configuration, and audit logs, it keeps no durable business state. Any working tables or caches are disposable by design. It reads the authoritative system of records, computes recommendations, and writes them back through a narrow, versioned API. The moment master‑data tables, workflow builders, or user‑management modules appear, what is on offer is a system of records masquerading as “AI”.
Third, the commercial model must align fees with the economic quality of the decisions delivered, not with head‑count, storage volume, or page views. Seat-based, page-view-based, or gigabyte-based pricing creates a perverse incentive to keep more people in the loop, proliferate dashboards, and hoard data, since vendor revenue then scales with organizational inefficiency. A genuine system of intelligence should, on the contrary, make most clerical seats disappear. Contracts should therefore index payment to decision throughput and/or to a measured share of the value created (e.g., risk-adjusted working-capital reductions, gross-margin uplift, or quality-of-service improvements at constant capital employed), with explicit caps to bound risk for both parties.
Taken together, these negative tests discard most “intelligent platforms” and let the practitioner focus his attention—and budget—on the few offers with a credible shot at improving supply-chain decisions.
A common misstep is upgrading a system of records simply because it “looks old”. Executives see IT progressing by leaps and bounds and infer the newer system must be much better. Furthermore, the look and feel of the latest versions proposed by vendors is usually more attractive than that of the old system. The inference is misleading. Progress in software technologies does not imply that systems of records themselves are progressing.
On the contrary, the core technologies and concepts underpinning systems of records have barely progressed since the 1980s. Current progress focuses almost exclusively on cutting development costs for technologies already heavily commoditized. Thus, for systems of records, the primary goal when upgrading should be to reduce fees paid to vendors. Similarly, the benefits of a better look and feel are usually overestimated: if the old system was snappy, fancier screens rarely make employees more productive. Highly productive interfaces are often perceived as hostile to newcomers precisely because they lack the hand-holding features that hinder productivity once employees know the system.
Collectively, without collusion or even coordination, vendors of systems of records have fostered a persistent state of confusion in the general public. In the early 2020s, although systems of records had been commoditized for over a decade, most companies were paying more for them than two decades prior. This state of affairs is broader than our supply-chain concerns, but supply-chain software may be the most telling example. This again illustrates the adversarial nature of supply chain challenges, where incentives are only partially aligned.
Thus, systems of records are indispensable—and tightly bounded. They are the firm’s authoritative ledgers, not its calculators or its strategists. Whenever they are asked to report heavily, optimize, or “plan”, resource contention, semantic drift, and accountability fog follow. The right architectural stance is austere: keep the ledger thin, central, and boring; push descriptive analytics to systems of reports; push decision-making to systems of intelligence; fence every interface with a narrow, versioned API; prune functional overlaps relentlessly, even if vendors resist.
The corollary is as managerial as it is technical. Treat the ERP (and its cousins) as sacred ground: it remembers facts and enforces workflows—nothing more. Every extra feature that creeps in—KPIs, forecasts, optimizers, “what‑ifs”—quietly taxes performance, muddies semantics, and diffuses responsibility. If something is computationally fat, probabilistic, or economically arbitrated, it belongs outside the ledger.
With the role and limits of systems of records made explicit, we can now turn to what lives inside them: data. The next section examines how those raw symbols become information, why semantics—not missing rows—is the true hard part, why “bad data” is the wrong scapegoat, and how to extract, document, and harden what the decision engines will need.
Data
In digital form, business data are by far the most important sources of information for supply chain decisions. These data are collected primarily through the company’s systems of records. Other sources exist, but they matter far less. From a supply chain perspective, the relevant information lies within this data; extracting that information proves harder than generally assumed. Furthermore, the data’s complexity is likewise underappreciated.
Data are critical because supply chains are not directly observable; one sees them only through the company’s records. One could argue that supply chains ceased to be directly observable—by the senses in any meaningful way—well before the advent of computers. Indeed, by the end of the 19^th^ century, the rise of the giant corporation brought armies of bookkeeping clerks. Thus, even then, the supply chain was seen only through ledgers kept by those clerks. Many objects in human knowledge, like supply chains, are not directly accessible to the senses; the consequences are substantial.
As noted earlier, many supply-chain authors either ignore software or treat it superficially. Data are no exception. In particular, data are usually conflated with the information they contain, as if transforming the former into the latter were a given. Nothing could be further from the truth, and this transformation alone often consumes more than half the technical effort in a supply chain initiative. Moreover, far from being mere software technicalities, this transformation requires a deep grasp of the supply chain challenge. As a result, it is profoundly misguided to think this task can be delegated to IT, or to any specialists who are not, first and foremost, supply chain specialists.
Against this backdrop, a stock explanation is wheeled out whenever a planning initiative stalls—the data are bad. Past failures—rarely investigated—are paraded as proof. The slogan is convenient: it absolves vendors and integrators, blames an amorphous scapegoat, and discourages any examination of the ledgers. In digitized supply chains, the transactional record is usually good enough; what is missing is the hard work of turning records into information—semantics clarified, provenance and authority established—and then into decisions. The sections that follow separate those steps and make their failure modes explicit.
Origin of the data
Most business data lives in systems of records, yet those systems collect it only incidentally. The point of systems of records is to digitize the firm’s information processes, making them more reliable and productive. Data are not the intended outcome; it is a by-product of those operations. This insight is simple, yet its consequences are frequently misunderstood.
Data are organized to mirror the system’s quirks. Such intricacies arise from subtle performance constraints, leftovers from aborted upgrades, outdated concepts from an earlier development era, or the routine compromises of software development.
To make this concrete, consider two commonplace examples showing how record structures reflect engineering trade-offs rather than analytical needs.
Decades ago, a well-meaning engineer—working on a still-popular
ERP—stored product returns alongside sales orders in the sales
table
with negative quantities. With that trick, summing sales over time
yields a figure “almost” net of returns. The design, however, has
several downsides and proves far less clever in practice. For service
quality, relying on net sales biases analysis toward overconfidence
about stockouts. If 10 units are sold in one week and 3 are returned the
next, at least 10 had to be in stock in the first week—not 7—to serve
demand. Conversely, if most returned units can be reconditioned and
resold47, then replenishing inventory right before season’s end is
likely to generate sizable overstocks, as stock will naturally benefit
from the intake of returns. More generally, returns are not sales; while
they may share fields (customer, product identifiers), conflating them
in one table forces columns that make sense for only one of the
two—e.g., a cause_of_return
field.
Consider a simple online store front-end. After each transaction,
stock-on-hand must be decremented to keep availability current for later
patrons. Updating stock-on-hand is a straightforward UPDATE
on the
table that holds those values; call this table stock
. However, as
virtually every team behind a moderately popular store has found48,
updating stock
after every transaction creates heavy contention in the
database—precisely around that table. Indeed, stock
is
sizable—typically tens or hundreds of thousands of rows—and each update
can trigger a cascade in the app’s front end. The usual fix is to add a
pending_updates
table that queues stock changes. On each transaction,
rows are added to pending_updates
instead of updating stock
. Then,
every minute or so, pending_updates
is purged while stock
is updated
in batch. As a result, stock-on-hand information is split across two
tables rather than one. The pattern is sound, but a supply-chain
specialist must decide whether ignoring pending_updates
—purely for
simplicity—remains acceptable.
Many other cases tell the same story: data in systems of records was never meant for data‑science work. For the engineers who wrote the software, it was at best a secondary concern, often an afterthought.
Corporate managers are frequently surprised by the time and effort required to extract the right information from business systems. Many attribute the overheads to the IT department’s supposed inefficiency and to overcomplicated tools. That impression is reinforced by IT specialists who hide behind technical opacity to justify costs and delays. This perspective, however, is mistaken. Data in systems of records is readily accessible and, assuming a mainstream database, the query tools are nothing short of excellent. Costs and delays are high because turning data into information is surprisingly difficult—especially when attempted by IT specialists lacking the field expertise the task requires.
Semantics of the data
A typical enterprise system of records stores tens of thousands of fields; the largest exceed one hundred thousand. The semantics of those fields—what they mean, the purpose they serve, the behaviors they trigger—are usually, at best, loosely documented. That documentation may come from the original vendor or, for in-house systems, from the company itself. Custom extensions—ubiquitous in enterprise software—further blur meaning. In practice, there is rarely more than a short paragraph per field, and too often it merely paraphrases the field name.
To see why semantics matters, consider a single field: “order date”. That one timestamp might denote checkout time, last edit, acknowledgment, or payment clearance. Each reading drives a different operational reality: it shifts measured lead times, alters service and fill-rate metrics, and changes how cash flows and liabilities are recognized. Ambiguity at this level quietly corrupts every downstream calculation.
In an ideal world, two things would hold: the builders would leave precise field-level documentation, and management would enforce day-to-day practices that keep meanings stable in operations. In practice, three forces break that ideal: age (systems and notes outlive their authors), field-level inconsistency (meanings drift across modules and versions), and operator-driven semantics (users repurpose fields to get work done). The next paragraphs examine each in turn.
By software standards, most systems of records are old—two or three decades are common. Early implementations rarely came with field-level documentation and, even when they did, the authors long ago moved on. The code has been patched, migrated, and extended so often that the original intent is unrecoverable from commit messages. Crucially, field semantics are inseparable from the behaviors wired to them, and those behaviors shift across versions and modules. The net effect is predictable: large swathes of meaning are simply lost, and reconstructing them amounts to reverse-engineering—a slow and expensive task. Outdated—or outright obsolete—notes are the bane of long-running software initiatives.
Loss is only half the story. New meanings also accrete informally as operations adapt. Teams invent workarounds, reinterpret labels, and attach new workflows to convenient fields. Over time these ad hoc practices become the system’s de facto contract, regardless of vendor intent or what the manual still says. This emergent, undocumented layer drives semantic drift: the “current meaning” of a field is whatever the data-entry crowd now uses it for—and it can differ by site, period, or role. Pinning this down requires observing practice, not just reading schemas.
In practice, users are structurally unaware of this drift. Meanings
accrete over years—often longer than anyone’s tenure in a given role—so
today’s interpretation simply feels “the way the system works.” If the
application exposes two near-synonyms such as Discontinuation Date
and
Phase-out Date
, and neither drives a visible automation, teams adopt a
convenient house meaning and move on. Whether that meaning matches the
vendor’s intent is immaterial; what matters is that the local crowd
converges on one reading and uses it consistently. Nobody will mistake
Volume
for “heel height,” but many will, quite sensibly, agree that
Discontinuation Date
marks the moment remaining stock is routed to
off-price liquidation. Given that most fields trigger no behavior at
all, divining the vendor’s semantics is wasted effort; ensuring everyone
inside the firm uses the same semantics is the real work.
Furthermore, field semantics are often row-dependent; not all values of a given field mean the same thing. We saw such an example above with “order quantity” referring to a purchase when positive and to a return when negative. More generally, a field’s semantics often depend on the value of another field (or several others). Documentation cannot be a mere dictionary of fields; meaning changes with context. To capture row-dependent semantics, documentation must cover the relevant business logic—again a costly, lengthy exercise, worsened by years of system evolution.
Ultimately, meaning rests with the data‑entry crowd—employees, clients, even suppliers—whatever the engineers once intended. For example, if Purchasing repurposes an obscure field—once unused and tied to arcane Incoterms—to encode a supplier’s transport route, then the semantics of that field become the transport route, not the original documentation. While repurposing a field is hardly good practice, it happens routinely. Moreover, as field semantics are tricky at the best of times, teams make genuine mistakes, attributing close-but-not-quite-right meanings to numerous fields.
Pinning down field semantics is hard—even “market-leading” systems prove surprisingly and invariably arcane. This work is not pedantry: without stable meanings, raw records cannot be turned into information a decision engine can use. Beyond the practical obstacles above, one more hurdle remains: we must be able to verify that the meanings we ascribe are in fact correct.
Even a modest supply-chain initiative typically requires documenting at least a hundred fields. For each field, half a page to a full page is usually needed to pin down the semantics that matter for decisions: definition; units and admissible values; provenance and authority; row-dependent interpretations; cross-field dependencies; and concrete examples. These notes should be grounded in the firm’s own data—simple distributions, ranges, and counts often surface what matters (for example, a typical hypermarket in our portfolio carries more than 100,000 SKUs, a fact that rules out algorithms whose cost grows superlinearly with the number of items). As a consequence, the semantic dossier for a “small” scope easily runs to a hundred pages or more. Thus, given the volume, it is unreasonable to expect all of it to prove factually correct. Errors—ranging from crass mistakes to subtle misunderstandings—will litter the documentation. Hence a method is needed to assess the validity of this knowledge and weed out the errors.
Extra reviews—even peer reviews—rarely catch every error. Indeed, once a misunderstanding is solidified in the documentation, a new element is usually needed to reveal it. That element must be real-world feedback from a decision-making process that draws, among other things, on this knowledge. If flawed knowledge yields sound decisions, the flaw is harmless and the issue moot. The fine print of this methodology—experimental optimization—will be detailed in a later chapter. For now, suffice it to say that semantic validity must ultimately be rooted in empirical feedback from the practical application of this knowledge while generating decisions.
Shenanigans in data
Data, like everything else in supply chain, bends to the incentives of its makers and collectors. As noted earlier, incentives often diverge from the company’s interests. This yields numerous quirks in the data—some minor, some consequential.
Consider a large manufacturer’s real-world mishap. The manufacturer was suffering from chronic overstocks. A glance at past orders revealed routinely exuberant buys—the main cause of the overstocking. To redress the situation, management chose a new supply chain vendor to fully automate replenishment, removing well-meaning but misguided judgment calls from the purchasing team. Pre-production tests showed the excessive orders vanished. This still left the possibility of overstock whenever a large, unexpected drop in demand occurred, but instant overstocks absent major market swings were eliminated. The solution looked promising.
Yet two months into production, excessive replenishment orders persisted. The result was puzzling: the orders bore no resemblance to the recipe’s output.
Closer inspection suggested a mundane cause: purchase orders were still keyed into the ERP by hand because bidirectional integration had been deferred—the IT team was overloaded. Historical data extraction from the ERP—a venerable four-decade-old system—was automated, but writing back the new orders was not. The decision engine produced orders unattended; clerks then retyped them into the ERP each day. This remaining manual hop became our prime suspect.
However, after days of careful assessment, it became clear that while manual data entry was involved, this step was performed very carefully, and that, under no circumstances, were employees intentionally modifying the quantities. Thus the new solution again became the primary suspect. Yet log inspection quickly showed the new solution had never generated those excessive orders. Dates and products matched; quantities did not. The bafflement deepened: day after day, the orders entered into the ERP the previous day exactly matched the solution’s output. Yet days—or even weeks—later, some reorder quantities suddenly ballooned, and the next inventory check confirmed the surplus.
Replenishment orders seemed to surge spontaneously, without rhythm or reason. So the vendor’s team studied those occurrences more closely. A surprising pattern emerged: the surges appeared chiefly in orders placed near quarter-end. Because lead times varied by product, the overstocking events were fairly evenly distributed, which had obscured the pattern. The calendar pattern was clearly human, but the actors remained a mystery.
Weeks of investigation led the vendor’s teams to the receiving docks, where merchandise was unloaded from supplier-chartered trucks. The receiving teams followed multiple quality-assurance steps. One step was to count the units and ensure they matched the ordered quantity. When a discrepancy arose, the ERP offered little to handle the edge case. In fact, the only official policy the ERP supported was to declare the shipment noncompliant and return it wholesale. Yet the receiving teams had learned this invariably incurred the production teams’ wrath. If goods had been ordered, they were needed; without them, production halted, infuriating colleagues.
Decades ago, to cope, the receiving teams established a backdoor into the ERP. To avoid returning a shipment for a single-unit discrepancy, the team used the backdoor feature to adjust the order quantity to the quantity received. The supplier was then paid the adjusted amount, reflecting what was actually received. For years, the setup worked, until the backdoor was essentially forgotten by all but the receiving team.
Over time, a few suppliers noticed that small over-shipments were accepted without so much as a phone call. This was unusual: clients typically returned the excess units or at least complained loudly. A few suppliers began exploiting the process. At quarter-end, if lagging their targets, they used the manufacturer as the variable of adjustment. At first, suppliers inflated orders only slightly, but the manufacturer’s silence soon emboldened them to submit far larger quantities.
Once the pattern was understood, identifying the misbehaving suppliers was easy. A series of phone calls to the shortlist of incriminated suppliers ended the explosive surges overnight.
Unexpectedly, though the pattern was baffling and obviously detrimental to supply chain performance, the data were fundamentally correct. The ordered quantities were real and accurate, as plainly demonstrated by the stock on hand. The error lay in the real policy governing replenishment—letting suppliers decide whatever they wanted—which the manufacturer had never envisioned.
While data shenanigans are not always as convoluted as the one above, unintentional adverse incentives routinely arise in any sizable supply chain, leading to all sorts of data oddities. We have already noted a common mistake in supply chain circles: treating data problems as matters to be left to IT specialists. The example above makes clear why this approach usually fails. Grasping supply chain data is first and foremost the province of supply chain specialists.
Shadow IT
In organizations, shadow IT denotes departmental systems that operate without central IT oversight—systems the main IT team may not even know exist.49 Shadow IT exists in virtually every large company, though prevalence varies greatly. From a supply chain perspective, shadow IT systems are predominantly loose systems of records and systems of reports, often mixing the two. Shadow IT matters for supply chain because the “system of records” portion often holds valuable data. For example, ERPs commonly omit some ordering constraints that apply to specific suppliers. As a result, to compose a valid replenishment order, the practitioner relies on his own shadow system containing a complete survey of the applicable constraints.
Spreadsheets are, by far, the most common technology for shadow IT. Any workflow that relies on employees interacting with a shared spreadsheet is almost certainly shadow IT. IT knows spreadsheets are used, and the software itself is approved. However, a spreadsheet-run system of records lacks what IT should expect of any supervised system: versioning, backups, audit trails, fine-grained permissions, etc.
Beyond spreadsheets, simple databases50 are also commonly used as part of shadow IT. These databases allow more polished systems of records, enforcing a modicum of type-level correctness (a value is a number, a date, etc.). Yet, like spreadsheets, they lack the supporting processes that ensure long-term maintainability (versioning, backups, etc.).
The motivation for shadow IT is straightforward: central IT is swamped for months—sometimes years—so business teams build what they cannot wait for. The paradox is that shadow IT is rarely born of excessive zeal for software; it is born of too little software culture. Software now touches almost every workflow, so most departmental problems are, in part, software problems. If the organization expects a single central IT group to solve all of them, overload is guaranteed.
The cultural deficit amplifies the backlog. Outside IT, teams often lack the vocabulary to state requirements, to rank them economically, or to estimate engineering effort. Pet projects are promoted as “quick wins,” essential needs are deferred because they look dull, and tickets arrive as solutions (“add a field”) rather than problems (“we need to stop double-billing returns”). Central IT, besieged by poorly framed requests, appears unresponsive; business units, forced to deliver, turn to spreadsheets and ad hoc databases. Friction becomes the norm and productivity erodes on both sides.
Central IT is also in a bind: employees are understandably reluctant to expose the full extent of their shadow initiatives, which often run with tacit managerial approval and plainly violate IT guidelines. As a result, the very people who could help regularize these tools discover them only when something breaks. Expecting central IT to “bring the data” is therefore unrealistic; by design, much of it is invisible.
The way out is straightforward, if not easy: replace shadow systems of records with small, maintainable CRUD apps, one by one. Prioritize ledgers first, reports later. Records are at risk of loss; reports can be recomputed once the underlying tables are sound. Concretely: take the live spreadsheet (or desktop DB), freeze it read-only, sketch a minimal schema that reflects the actual workflow, stand up a narrow CRUD service with versioning, backups, audit logs, and permissions, backfill the data, redirect users, then retire the old file. Rinse and repeat. Once the data sits in a production-grade store, rewriting the surrounding reports is cheap and fast.
Supply-chain specialists must own this remediation. Software is an auxiliary science of the practice, not an outsourcing target. Central IT should provide the guardrails (runtime, security, backups), but the scope, sequencing, and semantics must be driven by those who understand the flow of goods.
Derived data
Business systems store more than facts. Alongside event records (sales, receipts, stock counts, price agreements, calendar constraints) sit large volumes of what I will call derived data: numbers created from other records—aggregations, forecasts, EOQs, safety stocks, classification flags, and the like. Derived data add no novel information; they are a re-expression of what is already in the ledger. A beginner expects every table and field to carry “new” data; in practice, the majority of tables in a mature application are artifacts. For clarity, facts data refers to the original records of the flow, and derived data refers to anything computed from them.
Not all derived data are equal. One class is properly motivated: working or temporary tables (materialized views, caches, denormalized summaries) introduced for performance or operational convenience. They exist to make the system responsive and are disposable because they can always be recomputed from facts data. The other class is epistemic decoration: planning artifacts saved into ledgers (forecasts, EOQs, “service levels”, ABC tags, etc.). These tables also contain no new information; they merely freeze the output of a recipe. When mixed with records, they blur accountability, multiply inconsistencies, and later mislead systems of reports and systems of intelligence.
Demand forecasts are a classic example of such artifacts. More generally, almost all planning data held by companies are nothing but derived data. Unfortunately, many practitioners mistake these artifacts for stepping stones toward a system of intelligence; in practice they are landmines.
Before turning to why planning artifacts are misleading, let us dispatch a benign subclass of derived data. Engineers often introduce “working” or temporary tables—materialized views, denormalized summaries, staging areas—solely to keep CRUD ledgers responsive or to decompose heavy logic. These tables are conveniences, not knowledge; they carry no new information beyond the underlying facts and can be recomputed at will. Practitioners should largely ignore them for decision-making.
Misguided practitioners are naturally drawn to artifacts such as sales forecasts, economic order quantities, safety stocks, and service levels because these aggregates feel “digestible”. These aggregates compress thousands of row-level events into a handful of numbers—a weekly sales total looks far more actionable than a page of transactions—so their informational density for the human reader is high. Yet this convenience should not be confused with novelty: these figures contain exactly what the underlying facts already implied—and nothing more. Treating such summaries as primary inputs to a decision engine mistakes ease of consumption for added information and betrays a misunderstanding of the task at hand.
First, derived data contain no novel information. They are deterministic functions of facts data. As long as the underlying facts can be exported, every forecast, EOQ, “service level,” or ABC tag can be recomputed outside the original system. Moreover, CRUD ledgers are culturally and technically hostile to heavy computation. Sophisticated analytics belong in environments built for them; recomputing those figures in a non-CRUD setting is both simpler and more reliable than bending the ledger to do the math.51
Put differently—and to connect the discussion back to information—derived data contain zero new bits. They are deterministic functions of the firm’s fact data; at best they are lossy compressions, convenient for a human reader. Treating such artifacts as primary inputs conflates representation with knowledge: any uncertainty they appear to resolve was already resolved—no more, no less—by the underlying records.
Why, then, do ledgers teem with these tables? CRUD systems are engineered to keep workflows snappy and auditable, not to perform heavy optimization. Vendors therefore materialize intermediate calculations—forecasts, EOQs, safety stocks, ABC tags—so screens load fast and so, a ledger can posture as “planning”. This precomputation spreads because it is cheap to store and easy to demo, not because it adds information.
Second, handling these artifacts is materially harder than working with primary (fact) data—often by an order of magnitude or more. Flow events are bounded by physics; artifacts are bounded only by storage—effectively abundant. In software products that conflate ledgers, reporting, and “planning”, derived artifacts proliferate until they dominate the schema. It is common to see derived tables outnumber fact tables ten-to-one. The consequence is predictable: more volume to move, reconcile, and govern, with little or no informational gain.
Third, the analytical artifacts embedded in CRUD apps (setting aside purely operational caches and materialized views) almost always reflect naive, legacy approaches that do not belong in modern practice. This is not an accident but a consequence of the CRUD engineering culture: ledgers are built to keep workflows snappy and auditable on top of relational databases, so vendors gravitate toward recipes that are (a) expressible in SQL, (b) cheap to precompute, and (c) easy to demo. Forecast tables, economic order quantities, “service levels”, and ABC tags tick all three boxes; pricing optionality, cross-SKU capital arbitration, or risk-adjusted portfolio decisions do not. The outcome is a steady accumulation of brittle numbers that freeze simplistic assumptions into the ledger, blur accountability (which figure is authoritative—the artifact or the facts that generated it?), and mislead downstream efforts by masquerading convenience as knowledge. Had the authors aimed at economically sound decision-making rather than screen speed and sales collateral, these artifacts would never have been mixed into the system of records in the first place.
Not every statement about the future is derived data. Some future-dated entries are themselves facts: decisions or commitments that bind the firm or its partners. For example, if the marketing department has formally approved a budget for an advertising campaign starting early next year—complete with timing, regional splits, and channel allocations—this entry is not computed from other records; it is a new constraint that changes cash flows and capacities regardless of what demand later reveals. The same applies to effective-dated price changes already agreed with a supplier, planned store openings or closures with firm dates, or a regulatory change whose enforcement date is fixed. Such items should be recorded as facts, with clear provenance, effective/expiry windows, scope, and revocation rules.
More broadly, the habit of treating derived artifacts as inputs to a system of intelligence reflects a planning-centric worldview that puts planning on a pedestal. Yet planning—and the artifacts it emits—is a second-class citizen in supply chain. What matters are the enacted decisions that commit scarce resources. Intermediate computations such as time-series forecasts, especially when they support far-off choices like next season’s replenishments, do not deserve equal billing: they are provisional and will be revised repeatedly before any cash or atoms move.
Systems of intelligence exist to issue decisions—purchase orders, allocations, prices, schedules—that commit scarce resources. While exploring the option space, the engine produces a trail of intermediate calculations (scores, simulations, scenario logs, provisional forecasts). These are derived data: they contain no new information beyond the inputs and the code. Useful for auditability and diagnostics, they are transient technical by-products, not master data. The only durable outputs are the decisions themselves, which must be written back to the system of records through a narrow interface. Those entries are not “artifacts”: they are new facts—commitments that the firm will execute and that will later materialize as receipts, movements, and sales.
Consequently, a system of intelligence should consume authoritative facts from systems of records (and possibly a few external feeds with clear provenance), not the frozen artifacts embedded in ledgers or reporting tools. If a summary is needed, it should be recomputed inside the engine from the underlying facts; ingesting precomputed artifacts merely imports someone else’s assumptions—and their information loss.
External data
By external data we mean signals that do not originate in, and cannot be reconstructed from, the company’s own systems of records. They must be obtained from outside—public websites, official releases, third-party sensor feeds, or commercial data brokers. If the bytes can be derived solely from internal ledgers, they are not external; they are internal or derived data. Exploiting external data can create value, but acquisition, cleaning, and upkeep costs are high and often erode returns. As a rule of thumb, exhaust the economic value of internal data first; the recurring exceptions are competitors’ posted prices and a few highly curated, industry-specific datasets.
The chief obstacle with external data is the cost of acquiring, cleaning, and keeping it current. Unlike the firm’s own ledgers, external feeds are irregular, partially structured, and seldom aligned with internal schemas or refresh cycles. Internal records tend to scale with the firm; external sources do not. When external data prove worthwhile, it usually falls into three recurring families: (1) competitive intelligence—rivals’ posted prices, assortments, and terms—which directly modulate demand and constrain pricing; (2) weather observations and short-range forecasts, which can inform near-term allocation and scheduling; and (3) social-media exhaust, attractive in theory but typically unrewarding in practice.
Each family brings distinct frictions: pairing and normalizing items across catalogs for competitive data; locality, horizon limits, and heavy volumes for weather; and noise, manipulation, and weak causal links for social media. The next subsections take them in that order, starting with competitive intelligence.
Competitive intelligence
Competitive intelligence, in the supply chain sense, means harvesting rivals’ public signals—posted prices, assortments (styles, colors), and terms (warranties, delivery promises)—and folding them back into our own decisions. It is first and foremost an external data problem: these signals do not live in our ledgers; they must be acquired, matched to our catalog, cleaned, and kept current.
Economically, the link is direct: price rations scarce goods and coordinates substitution. Small price moves—ours or a rival’s—shift both total volume and market share, sometimes abruptly once retaliation begins. Thus, to allocate capital and inventory rationally, the supply chain must track competitors’ prices, assortments, and terms and let these signals inform its decisions.
Before the late 1990s, gathering competitors’ posted prices was tedious clerical work. The spread of e‑commerce in the 2000s made prices readily accessible on public sites. Even professional goods once hidden behind requests for quotes now surface, because a single reseller can publish the figure. Consequently, a cottage industry of web‑scraping experts—and tools to parse messy pages while impersonating human visitors52—emerged during the 2000s.
Today, companies can cheaply buy extensive datasets, refreshed several times a day, that map competitors’ assortments. Exploiting such data is still laborious; several hurdles remain.
Competitive‑intelligence data are far noisier than the company’s internal records. Web scrapers lean on many heuristics, and these can occasionally misread a value53. Even with high overall quality, wrong values must still be detected and neutralized.
In most markets our catalog does not map 1-to-1 to a rival’s. Even when a competitor sells the “same” article, its SKU, pack size, color set, or included accessories differ. Pairing is therefore a matching problem, not a lookup. A workable approach proceeds in three steps: (1) generate candidate pairs using cheap signals (shared manufacturer references, GTIN/EAN/UPC, normalized titles, salient attributes such as size/volume/power rating, and—when available—image embeddings); (2) score each candidate with multiple independent tests—text similarity, attribute compatibility, price-per-unit coherence, and bundle/variant rules—to produce a confidence score; (3) retain the full matching graph (one-to-many and many-to-one) with timestamps rather than a single “best” link, because assortments drift and the correct pairing can change over time. Modern language and vision models make step (1) easier, but they do not remove the need for steps (2) and (3). Without an auditable pairing layer—complete with versioned rules and confidence thresholds—downstream price comparisons will be numerically precise and economically wrong.
Data volume is not the only concern; persistence and provenance are. Competitive feeds are not “one row per SKU per rival per day”. To remain auditable and economically useful, the firm must retain: (i) every raw observation (timestamped URL, page or API snapshot, and the parser version used), (ii) the full matching graph with per-edge scores and timestamped rule versions, (iii) normalization rules and unit conversions as they were at the time, and (iv) every detected change (price, availability, variant set, bundle composition). Prices can move several times a day; marketplaces expose multiple seller-specific offers; variants and accessories appear and vanish. Collapsing this volatility to a single “daily price” discards exactly the signal we need for allocation and pricing.
Practically, the competitive corpus often exceeds the firm’s own transaction log by 10–100 $$\times$$ in rows, and by far more in bytes if raw HTML/JSON snapshots are archived for traceability.54 Modern hardware digests the volume; the real cost lies in the plumbing: ingestion pipelines resilient to site changes, anomaly detectors to neutralize scraper glitches, periodic re-matching jobs as assortments drift, and quality assurance flags that quarantine dubious observations without halting the feed. Neglect any of these and the dataset will be numerically dense yet epistemically thin.
Overall, competitive intelligence is consequential and affordable enough to merit inclusion in most supply chains. However, if a project can launch without competitive data, add that feed later, once the core recipe is live.
Weather data
Many categories—apparel, beverages, home and garden, automotive accessories—are weather-sensitive. Because weather moves demand, it is tempting to let forecasts steer the supply chain. The right question, however, is not whether to use weather data but when it truly informs a decision. Two practical axes determine the matter: the decision horizon (hours, days, weeks) and the spatial granularity at which the firm acts (store, city, region). A forecast carries useful information only when its horizon and locality match the decision’s. In practice, two hard constraints limit the payoff of weather-powered decisions.
Horizon: beyond roughly ten days, modern forecasts regress toward seasonal averages. From the empirical rules of the early 20^th^ century to today’s high-resolution satellite imagery, steady progress has improved forecasts, yet they remain a resolutely short-term affair. Consequently, choices with a look-ahead beyond ten days are no better informed by weather feeds than by plain seasonality measured on the firm’s own history.
Locality: Weather is intensely local. Conditions can diverge sharply across places separated by only a few kilometers—or even a few hours—while large firms sell into and operate across hundreds of thousands—or even millions—of square kilometers. A weather signal helps only if it is available—and mapped to our assets and customers—at the same spatial granularity at which we can act (store, route, city, or region) and within the relevant lead time. Achieving this match requires high-resolution grids and dense historical reanalyses to calibrate the statistical link between weather and demand. Covering the next ten days plus several years of history quickly yields datasets tens of gigabytes in size—often an order of magnitude larger than the firm’s own transaction log. Modern hardware digests the volume, but the plumbing is non-trivial. When decisions are taken at a coarser level (e.g., regional allocations), fine-scale signals largely average out, shrinking the payoff.
Furthermore, weather is not a one-dimensional phenomenon. Typical parameters include temperature, humidity, wind speed and direction, precipitation, pressure, visibility, and cloud cover. Depending on the goods, some parameters prove irrelevant, yet identifying them still takes work. For electricity demand, temperature often suffices, whereas other products hinge on a different mix of weather factors.
Finally, asymmetric risks in a given supply chain decision can undermine the benefits of weather-enriched forecasts. Let’s consider women’s razors. In Western countries, those products follow a sharp quasi-seasonality: demand surges in the week just before the first hot spring weekend. Otherwise, demand stays low for the rest of the year. For a general-merchandise store, the wiser move is to avoid last-minute, just-in-time replenishment the moment a hot weekend looms. If the forecast is wrong, replenishment happens too late, and the store has missed its once-a-year opportunity to sell those products. A sound, risk‑adjusted choice is to replenish in early spring. There will almost certainly be a first hot spring weekend; the uncertainty is its precise date. Keeping a pack of small, nonperishable products on the shelf a few extra weeks is far cheaper than missing the surge because of a bad forecast.
Weather helps only where its signal survives the two hard constraints already noted: horizon (roughly ten days) and locality (the unit at which the firm can act). Outside that window, the physics of forecasting and the plumbing costs dominate, while seasonal patterns and risk-adjusted policies carry more weight. Tooling may soften these frictions later, but for now, weather is an adjunct—not a foundation—for planning.
Social media data
During the 2010s, social-media data fascinated executives more than any other external source. As millions of consumers began voicing desires and opinions on social media, a company able to continuously refine its offering on that information could outcompete its rivals. Instead of relying on sales data observed with weeks of delay, the hope was to capture market sentiment in real time. Moreover, social media would hold information beyond the company’s current offering, providing clues to guide new-product introduction.
Yet social-media data never delivered the riches vendors promised; the expected benefits did not materialize. Indeed, social-media data are very challenging to exploit because it is extremely—and often maliciously—noisy. Sock-puppet accounts abound; genuine users troll or manipulate; influencers wax and wane. Inflated audience figures—views or subscribers—are commonplace. Social-media platforms have little incentive to purge click-farm traffic55 that inflates those figures.
Even if quality were higher, matching posts to precise parts of the company’s offer—at scale and automatically—remains onerous. While it is obvious that public feedback helps design or engineering teams by surfacing subtle flaws—or, conversely, areas for improvement—supply-chain feedback is mostly unhelpful. The company does not need social networks to learn that recent shipments were delayed; its systems of records—and carrier integrations—already provide that information.
Finally, processing costs remain consequential. Unlike similarly large datasets such as weather data, which are entirely numerical, social-media data are largely unstructured, with text and images representing almost all of it. Recent progress in multimodal LLMs (large language models) eases processing, yet running them at scale remains costly and slow.
Among external sources, social-media data exhibit the worst cost-to-benefit ratio. Extraction is arduous, and the insights are rarely actionable from a supply-chain perspective. In practice, firms should skip social‑media data for supply‑chain work until every other source is exhausted.
Mundane knowledge
A common objection to mechanizing decision-making is that people know many things that are not in the company’s systems of records. However, while this statement is a truism—the warehouse staff knows the color of the door handles, knowledge almost certainly absent from the system—it does not follow that near-complete automation is impossible. On the contrary, in the public and in corporate circles, a widespread mysticism endures about keeping people in the loop, as if the human mind possessed preternatural qualities no computer could emulate.
Let us start by clarifying the statement people know things with respect to information versus knowledge. While employees are certainly knowledgeable about the processes involved in the flow of goods, they are not necessarily highly informed, especially about the fine print of the flow’s ever-changing state. The human mind cannot reliably remember thousands of daily inventory positions, thousands of customer orders, hundreds of pending supplier deliveries. Even rudimentary computers vastly exceed the human mind’s capacity—both in reliability and in speed—to store and retrieve information.
Today, almost all remaining manual decisions in supply chains involve a planner who relies solely on information shown by company systems. When systems of records fall short, employees reach for spreadsheets—their shadow IT safety net—to assemble the missing information. This widespread practice illustrates that supply chains are de facto dependent on digital information. Whether this information is neatly integrated into well-engineered systems of records or haphazardly spread across a loose collection of spreadsheets, the case remains: for all intents and purposes, the information exists only in digital form.
Recall the distinction drawn earlier in the Knowledge section between special and mundane knowledge. Special knowledge—formulas, source code, contractual rules—already inhabits the written realm and thus is a natural feedstock for mechanization. Mundane knowledge, by contrast, lingers in tacit practices: knowing that carrier X never shows up after 4 p.m., that truck “Bessie” must be coaxed into first gear, or that a given supplier forgets the customs paperwork every second Tuesday. What skeptics of full automation really fear is that this fuzzy remainder will disappear once people are “out of the loop.”
The cure is not to renounce automation but to practice progressive codification. Whenever an operator relies on an unwritten rule that alters a material decision, the rule must be surfaced, written down, and placed under version control exactly like any other piece of productive code. Over time, the organization assembles a living, peer-reviewed corpus—a supply-chain playbook—from which engineers can harvest deterministic rules, statistical priors, or training labels for the forthcoming system of intelligence.
Cultivating such a playbook demands a company‑wide written culture. Meeting minutes, post-mortems, simulation notebooks, and corrective SQL snippets should be treated as first-class artifacts, stored in a shared repository for search, review, and refactoring. Written words decay slowly; oral traditions evaporate overnight. The sooner a firm embraces this discipline, the faster it converges toward fully informed, machine‑executed decisions.
Therefore, the information-based objection to automation does not stand. However, this does not address knowledge itself—and the intelligence needed to wield it. This objection can be restated: given all the necessary information, we still do not know how to formalize the transformation of information into decisions. Employees can do it, but they cannot say how—at least not precisely enough to permit reliable replication by a computer. This objection has merit and will be revisited later in the chapter Intelligence.
Bad data
Over the last four decades, supply-chain planning initiatives have been an almost unbroken string of failures, as demonstrated by the absence of production-grade unattended decision-making processes in most companies. While this statement will be revisited in greater detail later, let us immediately note that bad data has become the favorite excuse. Software vendors, software integrators, and data scientists alike invariably attribute their lack of results to inadequate data. Yet, this explanation does not stand up to closer scrutiny.
Since the late 1990s, all sizable supply chains have been digitized: every purchase, every shipment, every sale, and every production run is logged in a system of records somewhere. This does not imply that those systems capture all relevant aspects of the physical flow, nor that they are free of limitations and other issues. Yet, for more than two decades, the core transactional history of large companies has been digital.
Empirically, transactional data held by systems of records almost invariably exhibit error rates as low as one faulty entry in a thousand. This accuracy is unsurprising. Maintaining a precise ledger is a matter of survival for companies: unpaid customer debts must be recovered, and supplier invoices must not be paid twice. Routinely failing at those basics either guarantees bankruptcy or, at best, an acquisition by a better-managed competitor, since those problems are largely fixable. Occasional clerical errors still slip through despite safeguards, but they are too rare to meaningfully affect supply-chain optimization.
Non-transactional data held by systems of records are typically lower in quality than transactional data, precisely because erroneous entries do not systematically trigger an economic sanction. However, in supply‑chain settings, these data are seldom outright garbage. Indeed, most large companies are disciplined enough that their records reflect daily operations. The main quality issue with non-transactional data is that entries may be missing if employees can evade the data-entry process. When the user experience is poor, skipping data entry may be the only practical way to stay productive. Yet those fields go unfilled precisely because they are not critical to keeping the flow running.
In a sense, the very fact that the company keeps operating proves the data are “good enough”. Decades ago, some divisions kept paper backups, but such practices have vanished from large supply chains. If data are not in the company’s systems, it does not exist. As a result, missing non-transactional data does not prevent supply-chain optimization; it merely limits its effectiveness. Furthermore, the same data limits constrain employees performing the task manually. Few situations exist where the information in people’s heads suffices for optimization. Instead, they rely on context-free policies that keep the flow running—possibly inefficiently—on scant data. Yet those policies are not exclusive to the human mind; software can reproduce them.
Let’s consider, for example, a retail network with a warehouse serving a few dozen franchised stores. Each day, the independently run stores send orders to the warehouse, expecting next‑day fulfillment. Yet, the total quantity ordered by the stores on a given day might exceed the stock on hand in the warehouse. In this situation, the default policy is simply “first‑come, first‑served”. This naive rule invites gaming: stores overorder whenever they sense a possible shortage, which only deepens the problem. A better policy would distribute the remaining stock across the network rather than granting it all to the earliest claimants. However, such refinement requires more data, i.e., more context.
These observations cast doubt on the popular claim that “bad data” sinks most supply‑chain projects. After all, most initiatives amount to flow planning that needs only transactional data—data that is both available and reliable. An alternative, more convincing explanation is that most technologists—software vendors, software integrators, in-house data scientists—fail to exploit the data already sitting in the company’s systems of records. Indeed, supply-chain books and software alike tend to dismiss the problem: data preparation is wholly ignored, implicitly treated as a task so straightforward that it does not deserve any specific methodologies or technologies.
Yet, as detailed in the previous section, systems of records were never designed for data science and are unlikely ever to be. Large companies often operate many—sometimes dozens—of functionally overlapping systems of records. Such a chaotic application landscape arises from mergers and acquisitions, from successive roll‑outs that never retire old systems, and from ever‑new functional demands. Thus, data preparation—especially pinning down field semantics—is invariably a huge challenge for any sizable company. That challenge can only be tackled by supply-chain specialists.
In short, “bad data” often serves as a convenient scapegoat for plain incompetence. While external sources often suffer from overwhelming quality problems, this is not the case for the systems of records operated by the company itself.
Intelligence
Intelligence is the elusive faculty that lets people articulate their wants and plan the actions needed to reach them. It is self-evident that supply chain, understood as mastery of optionality, demands substantial intelligence to execute profitably. Yet a supply chain run crudely saddles the company with constant, avoidable overhead—costs sharper resource allocation would spare.
In an unhampered market, firms managed intelligently are rewarded, while those that are not are weeded out. The market is remarkable for delivering capabilities its participants do not explicitly understand. Earlier we saw the market deliver prices; now we see it delivers intelligent management as well. Those capabilities are obtained only indirectly—an aggregate property of the market as a whole. An executive must still rely on his own intelligence to improve his firm—specifically its supply chain.
Understanding intelligence is crucial to supply chain because much of it can now be engineered and delegated to computing hardware. In practice, for well over a decade, at-scale, unattended automation of the mundane operational decisions that keep goods flowing has been within reach. This is not a claim that general intelligence is “solved”; rather, it is an observation that specialized, engineered intelligence already delivers superhuman consistency on narrow tasks and large productivity gains. Today’s ubiquitous route-planning apps offer quiet yet powerful evidence.
The architectural form of this engineered capability is the system of intelligence. Before contrasting general and specialized intelligence, let us examine these systems on their own: how they arise in a modern supply chain, how they turn raw records into profitable unattended decisions, and why their design principles diverge sharply from the CRUD applications dominating enterprise software.
Systems of intelligence
Decision-making is mechanized through systems of intelligence. Unlike systems of records or systems of reports, systems of intelligence are engineered as replacements for human operators, not for man–machine interaction. The purpose of a system of intelligence is to generate the most profitable decisions, and to have those decisions made in a fully unattended manner.56 For most companies, relentless automation has been the cornerstone of profitability. Systems of intelligence extend this principle to tasks traditionally delegated to the white-collar workforce. Our ability to engineer a system of intelligence depends squarely on our ability to mechanize intelligence, whether general or specialized.
Although every corporate function—marketing, finance, HR, and so on—will ultimately have systems of intelligence of its own, supply chain is one of the most promising segments. Any sizable supply chain demands tens of thousands—sometimes millions—of decisions each day. Absent systems of intelligence, dozens or even hundreds of clerks are needed just to keep goods moving. By the early 2020s, many retailers and manufacturers employed more white-collar staff updating spreadsheets than blue-collar staff handling the goods themselves. Blue-collar roles have been aggressively mechanized for two centuries—a process still underway. By contrast, nearly all productivity gains from enterprise software in the past five decades have come from systems of records alone. White-collar headcounts have hardly fallen; rising supply chain complexity has absorbed the gains.
Yet the first inventory control systems, deployed in the 1970s, were expected to deliver extensive decision-making automation57. Those systems almost invariably relied on the safety stock paradigm, or on minor variants of it: for every SKU, the buffer is sized from a demand forecast, a lead‑time input, and a target service level. The method assumes demand uncertainty follows a normal distribution. The original expectation was that such a system could automatically generate inventory decisions, under modest supervision from employees to update parameters once in a while. The expectation proved misplaced; the systems remained intensely labor-intensive.
Yet in the 21^st^ century, after a few decades of stalemate, the practical automation of mundane supply chain decisions became possible through systems expressly engineered as systems of intelligence. For supply chain practitioners, the importance of those systems cannot be overstated: the modern practice of supply chain is, for the most part, a specialized software undertaking. It amounts to engineering systems of intelligence dedicated to supply chain. This fundamental insight is routinely missed by supply chain authors, consultants, and academics alike. As a result, the traditional paradigm fails to turn the supply chain practice into a productive asset.
A productive asset
What if we train employees and they leave? Indeed, but what if we don’t and they stay? (old business saying)
Before software, firms had to hire and keep clerks for every routine decision, no matter how small, mundane, or repetitive. As decisions multiplied, so did clerks and their managers. A company could narrow its focus to cut this headcount, but only by shrinking its offering. In this view, the staff is a pure operating expense (OPEX): the staff’s workdays are spent and leave no lasting asset once a decision is made.
Even with training and refined organization, large workforces have long since reached their competence ceiling. For instance, APICS58 (American Production and Inventory Control Society) was formed in 1957; its teachings have been stable for decades. Despite what many authors of the consulting variety still advocate, it is now unreasonable to expect that revamping the supply chain processes or organization, while remaining anchored in the pre-software paradigm, will yield anything but thin returns.
The fundamental proposition behind a system of intelligence is to turn those operational expenditures into capital expenditures. Even the man‑days invested in building the system improve the system itself. Operating and maintaining it still costs money, yet—as software—those outlays are typically a small fraction of the payroll required to achieve the same results manually. Thus, through the system of intelligence itself, the practice of supply chain is reified as a productive asset. This asset is productive because it generates ongoing returns for the company, through the decisions it automates, well beyond its operating costs.
Unlike a clerical workforce that can only be trained to a point, there is no a priori upper bound on supply chain performance that can be expected from a system of intelligence. First, it becomes possible to deliver what amounts to definitive fixes to certain inefficiencies or mistakes. Indeed, no matter how well-trained and how diligent the managers might be, any manual process is bound to suffer occasional clerical errors. Moreover, introducing further redundancies in manual processes—such as extra validations by peers or by the hierarchy—is usually counterproductive, as it slows the flow and adds hurdles and costs. Second, every decision can be further refined through more data, more accurate predictors, more effective optimizers …with no built‑in ceiling on their sophistication. Software can already package what would otherwise require lifetimes of accumulated insights and skills, and this has been the case for decades59.
As the maintenance of a system of intelligence is a problem of general intelligence, no system is yet fully autonomous. Since the early 2010s, however, a growing number of companies have run systems of intelligence that keep the mundane execution of their supply chains largely unattended60. Numerical recipes remain untouched for weeks, if not months.
Muddy emergence
Since the 1970s, supply chain software vendors have repeatedly claimed that their products would automate planning; in practice, these promises did not materialize—the systems shipped did not make unattended decisions, and for most vendors this remains true in the mid-2020s.61 Genuine systems of intelligence emerged only in the late 20^th^ century, first at scale in quantitative finance as algorithmic trading platforms—often referred to as “trading algorithms” or “automated trading systems”. Starting in the 1980s, financial institutions employed algorithms to automate trading decisions, effectively replacing human traders in various market activities. These systems were designed to analyze vast amounts of market data, identify profitable trading opportunities, and execute transactions at speeds unattainable by humans. The success of these systems highlighted the transformative potential of mechanizing intelligence, setting a precedent for other industries.
However, in other industries—most notably in supply chain—the confusion between systems of records and systems of intelligence has persisted since the rise of enterprise software in the 1970s, with software vendors—aided by market analysts—playing a critical role in perpetuating it. In a nutshell, systems of records are easy to design but offer limited returns; systems of intelligence are hard to design but have vast potential. As a result, most vendors frame their systems of records as systems of intelligence to inflate their prospects’ willingness to pay. Crude as it is, this marketing technique remains effective, with companies in the 2020s still paying high fees for technologies (CRUD apps) commoditized over a decade ago. Many executives remain convinced that systems of records contribute to the quality of their business decisions, despite several decades of contrary empirical evidence.
The value a system of records can deliver is strictly capped, no matter the quality of the solution. Indeed, once the system achieves near-perfect reliability and complete functional coverage, there are no further improvements to be made. For many processes—invoicing, inventorying, time tracking …—this endgame was reached more than a decade ago. As having the best accountant in the world makes no difference compared with having a great accountant, the greatest system of records makes no difference compared with a great one.
Enterprise software vendors have always been painfully aware of the inherent limitations of systems of records. Since the beginning, vendors have focused their discourse on the “better decisions” companies would make thanks to their software products. For example, in supply chain settings, vendors invariably claim that their solutions lead to lower stock levels and fewer stockouts, even when the software in fact entirely delegates inventory decisions to the client company’s staff—as is invariably the case with ERPs. Vendors selling systems of reports would later follow a similar approach, touting decisions that were, at best, only indirectly derived from their reports.
The crux of the technological challenge is that neither a system of records nor a system of reports can practically be extended into a system of intelligence. In defense of the early software vendors of the 1970s, the reasons weren’t readily apparent at the time, and thus the vendors themselves most likely believed that their early systems of records would, in time, encompass decision-making processes. Yet by the end of the 1990s, dozens of notable failed attempts had demonstrated that such a transition would not take place.
Culture and incentives
Automating supply-chain decision-making did not arise from the enterprise-software pioneers that grew into giants in the 1990s. This is doubly surprising: those vendors not only surveyed supply chains extensively—through their systems of records—but also relentlessly promoted their capacity to improve those very processes. Understanding why those vendors failed to deliver systems of intelligence for supply chain matters, because companies modernizing their supply chains will face the same problems.
One might argue that late 20^th^ century software was lacking and vendors were overly optimistic; that proposition is questionable. Indeed, the same vendors—primarily ERP and CRM vendors—still had not pivoted a decade after it became clear that some companies had mechanized supply-chain decision-making; they lacked designs that would make such automation possible. Thus, while technological barriers existed, they were not the primary hurdle.
The first problem is incentives. Vendors selling systems of records generally align fees with user counts; even fees not tied to headcount—e.g., support—track it in practice. As a result, vendors are incentivized to add traits that increase headcount and disincentivized from steering products toward reducing it. Systems of intelligence, whose explicit aim is to automate decision-making, run directly counter to that agenda.
Thus it is no surprise that vendors of systems of records—often highly profitable—have little reason to fund technologies that could erode today’s cash cows. If economics teaches one lesson, it is that incentives matter. If the organization’s profitability depends on a breakthrough not occurring, visionaries within it are inconsequential62.
The second problem is engineering culture. As noted, systems of records are developed almost exclusively along the CRUD paradigm. Compared to alternatives, CRUD demands far less engineering skill to maintain and extend an app. This has become a cornerstone of workforce management for vendors of systems of records. As a result, they can run larger teams while reducing dependence on engineering talent. By contrast, systems of intelligence invariably involve hard engineering challenges, scarcely comparable to CRUD trivialities. Unsurprisingly, a workforce that has dealt almost exclusively with “easy” problems—often for a decade or more—is reluctant to tackle genuinely difficult ones. As anecdotal evidence, the near-totality of computer-science breakthroughs in the last two decades came from consumer software companies63, while notable contributions from enterprise vendors of systems of records are hard to name.
Therefore, when rolling out a system of intelligence, a company should avoid vendors specializing in systems of records. Because systems of records are deployed first, managers are invariably tempted to turn to the incumbent vendor that has already proved trustworthy. Unfortunately, this trust is misplaced: the traits that make a vendor succeed at systems of records are wholly counterproductive for systems of intelligence.
Conflicting design requirements
Beyond incentives and culture, attempts to extend systems of records to encompass decision-making are largely doomed by conflicting design requirements. Performance, reliability, and maintainability matter to both camps, but what those notions entail varies drastically.
As we later explore technologies for supply-chain systems of intelligence, it will be clear that such a system cannot meaningfully grow out of a CRUD application and is best developed separately. Even basic considerations reveal how wide the gap is between systems of records and systems of intelligence.
Performance
A system of records must respond at low latency: whatever the input or screen, actions should finish within milliseconds—ideally under 100 ms—to remain imperceptible; beyond that, productivity degrades across mundane operations. Although 100 ms sounds fast, queuing ensures that a system averaging that figure will still inflict multi-second delays—latencies users notice and resent.
Therefore, in a system of records, every operation must finish in (effectively) constant time within a fixed resource budget. Because all operations compete for the same transactional core, any “fat” operation will eventually starve the system and degrade latencies across many operations. With thousands of features, even one exception is risky. Today, most companies with “slow” business systems suffer the predictable consequences of fat operations that should never have been implemented in the system of records.
It is easy to introduce such an expensive operation inadvertently. For example, summing an unbounded number of lines—e.g., computing a one-year moving-average revenue for a product category—already qualifies as problematic. Yet it is a basic piece of descriptive statistics. Readers versed in algorithms may object that most descriptive statistics can be maintained and retrieved within a constant resource envelope. While true, such sophistication is impractical64. First, CRUD’s virtue is that it avoids requiring algorithmic expertise. Second, at scale across thousands of features, performance woes creep in. Avoiding these operations altogether is the only scalable safeguard.
A system of intelligence faces a very different performance landscape. Consider deciding when and how much to order from an overseas supplier. Although the decision is time-sensitive, spending minutes—or even hours—on an optimization that yields a better order is sensible. Imposing a 100 ms cap on a decision that will take weeks to unfold makes no sense. The point of systems of intelligence is to generate the best decisions the company can afford to engineer. So long as computational spend is dwarfed by marginal gains, expending more is reasonable.
Reliability
Reliability means something very different for systems of records than for systems of intelligence. If a system of records goes down, any dependent process immediately stalls. For example, picking stock may not proceed if the inventory system is down. Thus, when engineering a system of records, it is critical to keep it operational at all times. Because CRUD applications isolate concerns, most bugs are local—they affect only one screen or process. Keeping the system of records running despite an identified issue is usually preferable to shutting it down until the bug is fixed. In fact, it should be possible to patch the system while operations are still in progress, as the resolution of a local problem should not be detrimental to the rest of the company. This capability—hot-patching—is generally considered technically difficult. However, the CRUD paradigm, via the relational database engine, largely delivers hot-patching. Thus, in practice, software engineers developing a system of records rarely require expertise in distributed computing to achieve near-perfect uptime.
By contrast, when a system of intelligence is suspected of being defective, prudence dictates shutting it down and diagnosing before resuming. Corrupted decisions can be hugely costly. For example, a vastly oversized order to an overseas supplier all but guarantees stock depreciation, if not an inventory write-off.
Consequently, in a system of intelligence, the economic quality of decisions matters more than uptime. When confidence drops, the engine should halt—deliberately and fast—based on explicit heuristics; this avoids expensive mistakes and is usually the most profitable choice. These halting heuristics are not the “alerts” and “exceptions” emitted when a system of records is stretched into a decision tool. CRUD designs cannot encode the fine print of nontrivial decisions, so edge cases proliferate: basic replenishment logic fails on launches, cannibalization, competitor promotions, and other routine shocks. Yet shutting the application down is unacceptable—edge cases are frequent enough that the tool would be down most of the time, and, being also the system of record, stopping it would paralyze operations. The tool therefore logs an exception and pushes the case to a human. A true system of intelligence does the opposite: the decision engine stops, and operations resume only after a programmatic fix—a change to the logic—has been deployed; hand-clearing cases is not an option.
Maintainability
The two flavors also differ sharply in their stance toward maintainability. In a system of records, once new entities are added, clients depend on them: data has been stored and processes now rely on it. Trouble starts when entities overlap and become unmaintainable. Overlaps create unfixable confusion—entities cannot be removed or even truncated. They also create conflicts when further developing the software. Similar situations occur when entities are ill-defined, ambiguous, or when they otherwise compromise the semantic integrity of the collection of entities. Thus, for a system of records, maintainability is chiefly a matter of carefully vetting entities and the behaviors overlaid on them.
Also, in a system of records, sheer codebase size is not, by itself, a maintainability problem. Many such systems hold thousands of entities across tens of thousands of tables. Larger codebases impose more overhead to extend the product, but do not jeopardize continuity. Furthermore, a hypothetical complete rewrite taking years is acceptable. While not desirable, it is not a prime concern: no situation calls for an immediate complete rewrite. If the business extends in a direction unforeseen by existing systems of records, it is acceptable to set up a new system to supplement them. Integration may prove challenging, but nowhere near as challenging as a complete rewrite of the older systems.
In contrast, a system of intelligence is maintainable if, once confronted with unprecedented conditions, it can be re-engineered quickly to resume operations. Such conditions can stem from a drastic strategic shift, a sudden market change, or any factor that fundamentally alters the flow. A system of intelligence should handle mundane variations, but a specialized intelligence cannot be expected to handle radical disruptions. That requires general intelligence, which is not readily available in computerized form.
For example, during the global panic of 2020, faced with a sharp decline in passenger traffic, some aviation MROs stopped repairing certain rotables (repairable components) and stored unserviceable items to preserve cash. Repairs may cost a third of the item’s price; with excess serviceable stock due to reduced traffic, delaying those expenditures makes sense. Before 2020, storing unserviceable items was not considered viable, and there was typically no process to opt out of repairs. All rotables eligible for repair were repaired by default. Unprecedented market conditions invalidated this policy. It would have been unreasonable to expect a system of intelligence engineered prior to 2020 to respond adequately to the lockdowns of that year. Systems in production had to be extensively modified to accommodate the new reality.
Because new conditions can demand a full rewrite of the decision logic, that rewrite must remain feasible. Practically, the team should draft a rough rewrite within hours and a polished one within days. Between such extremes, when a halting heuristic triggers, it should be possible to investigate and clear the case within hours—possibly less than an hour. This implies an engineer must always be capable of jumping into the core logic to assess and, if necessary, fix problems on the spot.
Therefore, the codebase of a system of intelligence should stay small and clear—ideally manageable by one engineer. Assume the existing logic will eventually encounter conditions that exceed its capacity to produce meaningful decisions. When this happens, there is no alternative but to resort to general intelligence from a capable employee to re-engineer logic for the new reality. In essence, maintainability for a system of intelligence is the capacity to rewrite, more or less completely, the system in the event of disruption.
No problem without solution
Systems of intelligence live or die by how they frame the business game they are asked to play. Even a flawless optimizer that pursues the wrong objective destroys value. Hence, before we let software reason on our behalf, we must first clarify exactly what it is supposed to reason about. This undertaking proves surprisingly difficult.
From primary school through university, students are presented with “problems” to be solved. Many—if not most—grades depend critically on students’ capacity to find solutions to those problems. The approach persists less because of pedagogical merit than because it gives institutions an easy, ostensibly objective way to grade students. However, this problem-solving mindset is deeply ingrained in our industrialized culture and has at least two notable consequences. These consequences matter greatly for supply chain and, more generally, for truly difficult undertakings.
First, there is a quasi-instinctive reluctance to think about “problems”. Academic institutions train and reward solution-finding, not the invention of problems. After two decades inside those institutions, this pattern becomes deeply ingrained in most of us. As a result, we often struggle to articulate a problem unless a solution is already in view. Thinking about solutions comes naturally—we have spent thousands of hours doing it—whereas thinking about the problems themselves is, for most, rarely practiced.
Consequently, major supply-chain problems are often ignored by practitioners and authors simply because no ready-made solution is at hand. Examples include price breaks, cannibalization, adversarial behavior, model uncertainty, and optimization under uncertainty—all consequential issues that today’s supply-chain discourse largely sidesteps. Companies address these problems indirectly through corporate policies—operating recipes that keep the business moving—but rarely state explicitly the precise problem each policy intends to solve. For a system of intelligence, this oversight is fatal: the code will faithfully optimize yesterday’s question while the market has already switched games.
Second, challenging the problem statement itself is instinctively perceived as challenging authority. Indeed, after two decades in academic institutions, every student knows that answering it doesn’t matter to a posed question guarantees a bad grade. Yet too often that answer is nonetheless the appropriate one. Careful selection of which problems to address is fundamental. Resources—time, computing power, people—are always limited, so we cannot tackle every conceivable problem. One of the entrepreneur’s most critical skills is wisely choosing which problems he decides to solve. This perspective radically differs from employees’, who are usually tasked not only with resolving a given problem but also with following a codified method embedded in corporate policies.
Both academia and business maintain incentives that discourage questioning the problem statement itself. In universities, publication, grant money, and student recruitment all depend on staying within officially sanctioned research fields. Scholars may pursue any “relevant” topic—provided it fits the canon of their discipline. Max Planck captured this dynamic in his Scientific Autobiography (1950): A new scientific truth does not triumph by convincing its opponents and making them see the light, but rather because its opponents eventually die and a new generation grows up that is familiar with it.
In companies, each problem’s solution is embodied in an org-chart and a budget line. Questioning the problem therefore threatens the status quo and creates managerial winners and losers. Politics are unavoidable in large firms and present a major hurdle to reframing the problem. Corporate history shows that organizations, like people, die—and refusing to confront new problems is a prime cause of decline.
Hence, the first deliverable in a system-of-intelligence project is not code but a clear, testable statement of the economic problem. Only after this statement survives adversarial scrutiny should optimization, forecasting, and data wrangling begin. With engineering and cultural groundwork in place, we can broaden the lens and revisit intelligence per se—first in its universal form, then in the specialized varieties that supply-chain work handles daily.
General intelligence
Systems of intelligence contain narrow, software-bound reasoning, whereas general intelligence is problem-agnostic and need not excel at any single task. Its hallmark is the ability to refine both the framing and the solving of problems. Riddles and tales—tools for developing intelligence in children and adults alike—have accompanied mankind since the dawn of time. A riddle is an exercise intended to sharpen the mind for its own sake. Tales play a similar role but leverage our emotional response to facilitate memorization.
Writing, invented over 5,000 years ago, is best seen as an amplifier of human general intelligence rather than its cause. It externalizes memory: perspectives, insights, and facts can be recorded, critiqued, and recombined long after their authors’ deaths. The printing press—and, much later, the web—lowered copying costs and expanded coordination, accelerating this feedback loop. Yet more storage does not by itself produce understanding or judgment. A gram-sized device can hold many lifetimes of literature and still be entirely unintelligent; memorization is not intelligence.
Intelligent animals such as orcas and chimps transmit cultural know-how—hunting or foraging tricks—to their young. They demonstrate notable intelligence, yet general intelligence eludes them. Contrary to past philosophical theories, no single trait appears uniquely human. Complex language, named individuals, humor, tools, formal logic, and theory of mind65 all exist in lesser form in the animal kingdom. The shift from intelligence to general intelligence is therefore a loose threshold—many mental faculties must cross a quantitative line before a qualitative leap occurs.
Until the early 2020s, deliberate self-improvement of general intelligence was strictly human and decidedly artisanal. The mind resists “engineering” in the sense of limitless, reproducible upgrades. Millennia of pedagogy have produced many techniques, yet no one expects breakthrough in this area, at least not in the foreseeable future66. A new language, linear algebra, or thermodynamics cannot be mastered overnight; disciplined study remains the only path.
By the 2010s, specialized intelligences could self-improve in narrow arenas—most famously Go67. The early 2020s brought a watershed: large language models whose largest instances display a discernible spark of general artificial intelligence68. Like the printing press or the web, these models expand humanity’s ability to refine its own intelligence.
With these foundations in place, we can clarify “general intelligence”. The next three subsections tackle the subject from different angles. First, we ask when a question becomes a general-intelligence problem—one whose option space defies formal bounds. Second, we look at clerical roles in supply-chain teams to show why they need specialized, not general, intelligence and are therefore ripe for automation. Third, we assess large language models—today’s strongest bid for engineered general intelligence—and sort real contributions from hype. Together, these lenses turn a philosophical idea into operational guidance for modern supply chains.
General intelligence problems
General-intelligence problems are inherently open-ended. Each solution merely deepens our understanding and forces a new formulation, which then demands a fresh solution. Breaking this cycle is never final: stopping points are partly arbitrary—a pragmatic truce set by today’s economics, tooling, and attention—until reality forces a new formulation. Such problems may feel subjectively rare because, once named, most “problems” are promptly wrapped in formal boundaries to make them tractable. However, nearly all business problems, in their naked form, are general intelligence problems.
Entrepreneurial profit-seeking is a clear example of a general-intelligence problem. A firm’s capital, time, and managerial attention are finite; the ways they can be deployed are effectively unbounded and continuously shifting. At every moment, the entrepreneur must decide whether to commit to the best currently known option or to expend resources searching for a better one—knowing the search itself consumes those same scarce resources. Because the option space shifts as technologies, competitors, and regulations change, no fixed formalization closes the problem. Entrepreneurship therefore belongs squarely to general intelligence. History shows entrepreneurs repeatedly redefining—or inventing—industries their predecessors never imagined.
By contrast, the vehicle-routing problem—choosing optimal routes for a fleet that must serve given customers—is the archetype of a bounded problem. Solving it may be hard69, but it does not require general intelligence. Any solution must obey the formal graph-theoretic statement of the task. The option set is finite—even if astronomically large—and every candidate shares the same shape: a permutation of stops (and timings) evaluated under a fixed cost model; no genuinely surprising move ever appears.
Every problem begins as a general-intelligence mess; turning it into a tractable “bounded” problem is itself an act of general intelligence70. The key question is whether the chosen formalization is adequate. Academia may catalog puzzles indefinitely, but businesses must pick one framing, because each resource can be allocated only once. We return to this methodological point later.
Conversely, any bounded problem can be lifted back into a general-intelligence problem by changing the frame that made it “bounded”—a move sometimes dismissed as “cheating” because it violates the original rules. In vehicle routing, the textbook problem assumes a fixed set of customers, service promises, fleet, and prices. The moment we renegotiate any of these—drop or defer remote deliveries, price them differently, add pickup lockers or a parcel carrier, switch vehicle types, widen delivery windows, or even make delivery unnecessary altogether (as streaming did to cassettes)—the original problem evaporates. We have not found a smarter route; we have changed the game. That reframing—altering constraints, commitments, or payoffs—is precisely the work of general intelligence.
Earlier chapters showed that several supply-chain challenges are genuine general-intelligence problems. Three stand out. (1) Selecting the economic model, which defines how actions translate into costs, revenues, risks, and constraints—the score we optimize. (2) Defining the option space: the set of admissible actions, timings, and quantities the firm is prepared to consider (a decision is a selection from this set, not the set itself). (3) Engineering the numerical recipes that, given data, choose decisions within that option space to maximize risk-adjusted returns while honoring constraints. Because adversarial behavior—by competitors, employees, suppliers, customers, or vendors—constantly pushes against boundaries, only general intelligence can tackle these issues. The story of suppliers exploiting a weak ERP to ship excess stock is one such subversion.
Much published supply-chain work remains rigidly attached to bounded formulations and leaves the general-intelligence design choices—objective, constraints, and option space—implicit. This does not make such problems intractable. Practically, surface those hidden choices explicitly—state the economic objective in monetary terms, enumerate admissible options, and define halting conditions—and then decompose the work into bounded subproblems a machine can solve. The next sections show concrete tactics for doing so and for iterating when reality pushes back. The program is necessarily incomplete—general-intelligence problems cannot be “closed”—yet it is sufficient to yield unattended, profitable decisions.
Clerical jobs in supply chain
Supply-chain challenges demand substantial general intelligence. A persistent misconception holds that routine decisions must be held to the same standard. This is not the case; they only require specialized intelligence—structured processes that can be engineered and handed to software.
The misconception stems from outdated practices. In many firms, clerks still decide, say, whether to raise a replenishment order. They do apply policy with intelligence, but general intelligence is scarcely involved—on closer inspection, almost none.
These policies are usually slight twists on 1970s-era practices71. Clerks seldom improve the fundamentals; they merely execute and maintain the rules. Even parameter upkeep—tweaking service levels, for instance—is a specialized, rule-bound task. General intelligence again plays no part.
While better reporting tools raised productivity, they did not improve the decisions themselves. Sparing employees the hunt for stock levels and sales history is useful, yet simply surfacing numbers does not, by itself, reveal how to turn them into profit.
Bluntly put, the 21st-century supply-chain clerk72 is the white-collar analogue of the 20^th^ century factory worker. Each day begins by opening an app—often a spreadsheet—and slogging through hundreds of rows. The cycle is endless: top performers become trainers or supervisors, perpetuating it. A desk does not, by itself, confer “intelligence”. Given today’s technology, these clerical roles will soon belong in history books—obsolete, to general relief.
Yet the accumulated know-how of those very clerks is anything but disposable. Their day-to-day heuristics—honed through years of wrestling with software vendor quirks, catalog edge-cases, and improvised workarounds—form the tacit domain knowledge that tomorrow’s systems of intelligence must absorb, formalize, and surpass. Therefore, automation is not a repudiation of their expertise; it is the vehicle that will preserve it, scale it, and finally liberate it from the drudgery that keeps it underleveraged today.
Before considering how routine tasks vanish, recall why “soft-skill” leadership loomed large in supply-chain teams. When planning was manual or semi-manual, leaders spent most of their energy coordinating armies of clerks—arbitrating priorities, smoothing friction between silos, and keeping throughput from stalling. That requirement falls away as soon as decisions are automated. Once an engineered—eventually artificial—intelligence carries the planning load, the coordination problem collapses; a handful of specialists can steer a very large flow of work.
The mandate changes. The leader’s job shifts from herding people to owning the machine that makes the decisions. That ownership has four parts: (1) state the economic objective in money terms and define the admissible option space; (2) guard decision quality—via audits, halting heuristics, and dual runs when needed; (3) steward semantics and data provenance, so inputs remain trustworthy; (4) recruit and protect a small, engineering-minded team able to modify the logic on short notice. Charisma and empathy still help, particularly when aligning finance, merchandising, and operations around the same objective, but they cannot substitute for technical stewardship. Firms that cling to clerical workflows will keep rewarding soft-skill theater even as margins erode—a theme we revisit when discussing manual forecasting.
Large language models
Large language models (LLMs) that emerged in the early 2020s mark a milestone in seven decades of rapid software progress. The most capable LLMs arguably show the first spark of general intelligence and thus merit close attention. They matter to supply-chain work—though many vendor-promoted “use cases” are mere buzzword-riding nonsense.
Conceptually, an LLM does one thing: it reads a sequence of symbols and predicts the next, appending it to the input. The input sequence is the prompt; the act of generating is sequence completion. Symbols could be single characters, but performance improves when the model operates on multi-character groups called tokens (vocabularies typically number $$\sim$$100,000). Text is tokenized before inference and detokenized after, restoring plain language73.
Internally, an LLM assigns probabilities to all possible next tokens. One may sample from this distribution instead of always taking the top choice, injecting randomness controlled by a temperature parameter (with temperature = 0 yielding deterministic output).
LLMs have a fixed context window—the maximum token count for prompt plus output. When that limit is exceeded, the oldest tokens are truncated and their information lost. Models now commonly hold over 100,000 tokens (roughly 100 pages of text). Techniques such as retrieval-augmented generation (RAG) extend practical reach, but lie outside this discussion.
Running a model (inference) is simple; creating one (training) is far harder. LLMs are deep learning networks with billions of parameters arranged as tensors74 and combined through mostly linear algebra operations.
Those corpora typically feature a trillion tokens or more. Training processes the corpus one token at a time and, each time, nudges the parameters75 toward a more accurate probabilistic prediction of the next token. After the public corpus is ingested, training repeats with a much smaller corpus intended to mold the model’s generative patterns toward those of a helpful assistant. This latter process is fine-tuning.
This summary also clarifies why the models are called pre-trained: unlike most machine-learning models, users need not bring their own dataset; the trainer supplies it.
LLMs embody an uncanny, non-human kind of intelligence. Software vendors oversell their capabilities, while many supply-chain experts—and entire organizations—downplay them, treating LLMs as a curiosity that leaves both theory and practice unchanged. Both stances are wrong. LLMs are neither magic nor negligible. They require basic prompting skills76 and proper instrumentation; without these, any model looks “dumb”. Their performance is highly task-dependent—subpar for arithmetic and strict numerics, yet superhuman for code de-minification, refactoring boilerplate, or rephrasing brittle business rules. A competent practitioner approaches them with prior beliefs about where they will shine and where they will fail. In that sense, prompting has joined spreadsheets and word processors as table stakes for white-collar work.
Contemporary models are also multimodal. Most frontier LLMs natively ingest images, and many now emit images as well, with the same integration spreading to audio—both speech recognition and synthesis—by shoehorning pixels or waveforms into pseudo-tokens. Yet outside these native modalities, the impressive “extra” powers that people see—web search, code execution, retrieval, optimizers, simulators—still come from an orchestrator layered on top of the model. The model emits a structured tool call; the orchestrator invokes the external tool; the result is fed back into context, and completion resumes.77 An uninstrumented LLM is a sharp autocomplete; the same weights, instrumented, behave like a competent analyst with a browser, calculator, and terminal. The intelligence in the weights is unchanged; the capabilities—and therefore the usefulness—are not.
Modern language models show a spark of general intelligence for two reasons: they can write code and they can follow chain-of-thought prompts. Their code falls short of an experienced engineer’s output, yet it already outperforms low-skill IT workers. The ability to write software is foundational, because self-improvement—the hallmark of general intelligence—requires it.
Chain-of-thought prompting coaxes a model to reason aloud. Merely appending “Let’s think step by step.”78 to a given prompt often boosts answer accuracy, but the extra reasoning tokens raise compute cost. The figure below illustrates the effect.
\setlength{\fboxsep}{6pt} \setlength{\fboxrule}{0.4pt}
Chain-of-thought (CoT) helps by forcing the model to unpack a task into smaller steps within its context window. The model retains no memory beyond that window; no hidden state carries from one answer to the next. CoT simply spends more tokens at inference time to surface intermediate reasoning. This usually improves reliability up to a point, with diminishing returns and hard limits imposed by the window. Recent systems increasingly perform this expansion under the hood and suppress the raw trace (“thinking” modes) to spare users long monologues. CoT is an inference-time budget to spend on harder questions—not a new capability in the model’s weights.
Despite their appeal, large language models typically consume a million-fold more computation than classic algorithms solving the same job. Hardware progress will lower per-token costs, yet these models are unlikely ever to rival purpose-built data-processing code. Reserve them for tasks plain algorithms cannot reach.
Language models matter to supply chain mainly in two ways: first, speeding up the writing and upkeep of numerical recipes and their documentation; second, extracting features from raw text. We revisit both later. Other applications are secondary.
Misuse of large language models
Since the early 2020s, many enterprise-software vendors have made bold—often nonsensical—claims about artificial intelligence (AI). They hide the details not to guard secrets but to mask how little technology lies beneath. Most of the hype centers on large language models (LLMs). Let us dismantle the main claims.
Our AI forecasts more accurately: Language models are ill-suited to time-series forecasting—or numerical work of any kind. Their performance, in both accuracy and computational cost, is abysmal compared with even basic statistical models, let alone top-ranked ones. Pretrained forecasting models—an LLM-style idea—are under active study. They may help elsewhere, but, as Chapter The Future shows, they are a dead end for supply-chain work.
Our AI crunches all your data: Per byte, language models are both expensive and slow, and will remain so for the foreseeable future. As a rule, processing megabytes of data through a language model is a nontrivial operation that takes minutes to complete. Large companies have terabytes of records in their business systems. Simply passing an entire corporate dataset through an LLM—even once—will remain prohibitively slow and costly for years. Cost aside, an LLM does only one thing—sequence completion. That tool is versatile, but forcing vast non-text datasets through it yields no obvious benefit.
Our AI talks to end-users: Because software vendors bill per seat, inefficiency sells: each extra clerk forced to fight the software means extra revenue. Bolting an LLM onto a screen to create a “chat” interface only worsens matters: conversation is slow and tiresome. Chatbots belong at help desks, not in high-throughput enterprise workflows; training people on a crisp interface beats sluggish dialogue every time.
In enterprise software, LLMs are today’s fashion—one in a long line soon to be replaced by the next. For forty years, CRUD systems have ruled, and that culture, dominant among vendors of systems of records, resists meaningful integration of new ideas, even when the ideas are sound.
Misconceptions about intelligence
Popular culture has carried the idea of artificial intelligence (AI) for decades, and much of today’s boardroom intuition still echoes 20^th^ century-century theories about what “intelligence” is supposed to look like. This matters because those intuitions leak directly into supply-chain decisions: they shape which projects are funded, which risks are feared, and which levers are left untouched. Before we mechanize decisions, we must clear away the myths that keep executives oscillating between paralysis and credulity.
In supply chain, the relevant question is not whether machines will “think like us,” but whether software can systematically make better bets about what moves, where, when, and how much, and do so unattended, with auditable economics. Misconceptions about AI push firms into two symmetric errors: alarmism, which delays profitable automation on the grounds that “AI is dangerous,” and anthropomorphism, which mistakes human-like conversation for decision-making ability and green-lights brittle toys.
The next pages examine three persistent misunderstandings—the “singularity,” the Turing-test view that human-sounding chat implies general intelligence, and the Moravec paradox, misread as evidence of machine limits—and recast them in operational terms. The standard for intelligence in this book is strictly economic: higher risk-adjusted returns from the same scarce resources. By that yardstick, most cinematic worries are irrelevant, and most demo-friendly chatbots are distractions.
The singularity
The singularity imagines an AI that self-improves so fast that it outstrips humanity, then either transcends or enslaves us. Countless action films riff on this theme, yet Hollywood’s grasp of AI is no better than its grasp of real-world ballistics. For supply-chain practice, this mythology is a non sequitur. The systems that decide reorders, prices, routes, or allocations are narrow, auditable programs with bounded option spaces, explicit objectives, and kill-switches; they do not “wake up”, they optimize under constraints. Refusing to mechanize replenishment or routing because “AI might go rogue” is like refusing forklifts for fear of robot uprisings: it trades certain, present waste for a speculative, distant hazard. The real risks are mundane and tractable—bad objectives, sloppy semantics, untested code—and so are the safeguards: dual-run, halting heuristics, code reviews, and economic audits. Existential-risk debates have no bearing on whether a firm should apply the best that computer science offers to the day-to-day allocation of scarce resources.
While the general public may get the impression that the technology leaped forward, essentially overnight, back in 202279, this is absolutely not the case. In reality, progress has been slow, incremental, and hard-won over the past seven decades. Furthermore, it is not one thing being improved, but dozens of largely unrelated fronts: better paradigms, better algorithms, better hardware, better datasets, better methodologies, and better codebases. All of these improvements have been generated by a diverse crowd of contributors. Thus, while it is possible to stumble upon something that exceeds human intelligence, the idea that such an entity would—as an event—elevate itself into a transcendent being is far-fetched.
Once we remove the “event” aspect of the emergence of artificial general intelligence, it becomes difficult to explain why this would ever present an extinction-level threat. Any malicious entity, or any entity operated by malicious people, will be countered by similarly capable entities that armies, intelligence agencies, companies, or even universities and hobbyists will develop. Artificial general intelligence will be a late addition to the already long list of technologies that can do immense damage when used for nefarious purposes.
Such an entity will be the product of an advanced industrial civilization and will rely on its complex—and fragile—supply chains. Semiconductor manufacturing already ranks among the most intricate industries. Conceivable “alternatives”—neuromorphic, photonic, quantum, or biosynthetic substrates—are, at best, lab-grade and themselves ride on the same semiconductor tooling, materials, and precision-manufacturing stack. In short, there is no industrial, non-silicon path to computation today that avoids those chains. The continued existence of such entities will depend on the active support of humanity for decades, if not centuries. People will have ample time and opportunity to address the inevitable problems that have accompanied the rise of every technology so far.
Doomsaying is ancient, and most experts and politicians join in for self-aggrandizement.
The Turing test
In 1950, British computer scientist Alan Turing devised his imitation game, later called the Turing test. The gist is that if a human, chatting through a text-only channel, cannot reliably tell human from machine, then the machine qualifies as generally intelligent.
The usual objections to the test dwell on consciousness and self-awareness; these are beside the point for supply chain. Our concern is not whether a machine “sounds human”, but whether it consistently makes choices that raise risk-adjusted returns under uncertainty.
Crucially, supply-chain intelligence is not self-evident to the naked eye. Humans have no sixth sense for whether “200 units next week” is slightly too high or too low. Even ex post, judging a decision requires counterfactuals—what would have happened had we acted differently—and horizons that match lead times. This evaluation is hard: outcomes are noisy, delayed, and entangled with thousands of parallel choices.
An engine that speaks poorly—or not at all—may therefore look “unintelligent” while quietly compounding value. Conversely, a fluent chatbot can pass Turing-style conversations yet steer the firm toward losses. In the 2010s, simple pattern-matching bots frequently fooled human judges; they were rhetorically convincing and operationally useless. The lesson endures: imitation of human style is orthogonal to decision quality.
A practical illustration: a good replenishment engine may place many tiny, oddly timed orders, reprice slow movers by a few cents daily, or divert one container to an unfashionable cross-dock. To a planner, these moves can feel fussy or even wrong; on the ledger they lower holding costs, reduce write-offs, and speed cash conversion. By contrast, a tool that “explains” itself beautifully while preserving legacy MOQs and batch calendars will earn applause in meetings and quietly bleed working capital.
The only meaningful test is economic. Call it an economic Turing test: under a dual run or via reasoned counterfactuals, with the same constraints and audit trail, does the system of intelligence deliver higher risk-adjusted operating profit than the incumbent process after all costs (inventory, logistics, markdowns, labor, capital) are charged? If yes, treat it as intelligent—no matter how taciturn its interface. If not, no amount of conversational polish redeems it.
Thus, fascinating though it remains, the imitation game is an inadequate criterion for our craft. Supply-chain intelligence systems should be judged by auditable economics over the appropriate time horizons, not by their ability to chat.
Moravec’s Paradox
The American computer scientist Hans Moravec observed in the 1980s that tasks that seem simple—grabbing a teapot and pouring tea—are among the hardest for artificial intelligence to replicate. In contrast, tasks once deemed marks of genius—such as matrix diagonalization—have proved straightforward for machines.
However, the paradox dissolves once its hidden premise is stated. The paradox assumes that animal mobility—and the planning behind it—is learned (nurture) rather than innate (nature). In other words, acquiring mobility is held to result from learning—one that has proved incredibly difficult to replicate artificially for the past four decades.
On closer examination, those capabilities involve little or no learning. Foals—and their equine cousins—stand and walk within hours. This is not “learning” but a quick calibration of a nearly finished motor system. The cognitive structures are nearly complete; the system needs only a slight nudge for the pathways to lock into place.
In humans, the innateness is less visible: compared with other mammals, our oversized heads force earlier birth, leaving motor circuits to finish developing outside the womb. What many perceive as the infant’s “learning” is cognitive development that would continue unimpeded if the infant remained in the womb for a few more months.
These largely innate circuits are evolutionary products, refined over roughly 800 million years and untold generations. No wonder machines struggle to match animal mobility. The challenge is akin to engineering a device that outperforms trees at converting sunlight into structural material, or ribosomes at assembling macromolecules. Outperforming evolution at games played for eons is brutally difficult.
Mobility rests on hundreds of tightly knit heuristics; even a snail can negotiate the three-dimensional maze of vegetation. Well-chosen heuristics can seem “intelligent” provided the action space is tightly bounded—as in the snail’s case.
In nature, mobility was a necessary evolutionary path to general intelligence: life could not leap from unicellular origins to general intelligence without first solving locomotion. Machines face no such constraint—so long as humans supply energy and materials while we solve the remaining hurdles (e.g., industrial substitutes for muscle fibers). Evolution solved these problems in ways that do not lend themselves to industrialization; we must solve them anew.
The implication for supply chain is easy to miss. The tasks that feel “simple” to humans—walking across a crowded aisle, grasping a box, recognizing a co-worker—are precisely those evolution has tuned for hundreds of millions of years. By contrast, large-scale, fast-moving resource allocation—ranking millions of micro-options with delayed, stochastic payoffs—is the opposite of what our brains evolved to do. Working memory is tiny, attention is fragile, and our intuitions about probability are notoriously poor. Machines invert this profile: they struggle with animal motion yet excel at enumerating vast option sets, scoring counterfactuals, and recomputing plans every minute. Thus the quip “AI can’t even fold laundry” is irrelevant to whether software can outclass planners at repricing 50,000 SKUs, rebalancing multi-echelon stocks, or rerouting containers mid-voyage under volatile lead times. In the Moravec sense, these are machine-native problems.
Two operational corollaries follow. First, do not condition decision automation on progress in robotics or conversational polish; the relevant yardstick remains the economic one introduced earlier: under the same constraints, does the engine deliver higher risk-adjusted profit than the incumbent process? Second, expect machines to overtake humans first where planners sweat most: combinatorial allocation, adversarial pricing, multi-echelon balancing, and dynamic rescheduling. None of this requires humanlike generality; specialized intelligence within well-defined boundaries suffices—exactly the habitat of supply chain engines.
In short, copying animal mobility is not a shortcut to general intelligence: it is a cluster of tough yet essentially specialized problems. Solving them may inspire, or occasionally inform, work on artificial general intelligence, but that remains a separate pursuit.
Specialized intelligence
A specialized problem is one whose options are bounded by explicit formal rules; solving it calls for specialized intelligence. The academic supply chain puzzles are typical examples, and each solution—whether a simple heuristic or a large hyperparametric model—counts as “intelligent” only by the quality of its results. Declaring a method “smarter” merely because it looks arcane is a mistake.
Supply chain becomes tractable precisely when its sprawling, general intelligence questions are carved into many bounded allocation problems. Each bounded problem fixes three elements: (1) the option set—what moves where, when, and how much may be chosen; (2) the constraints—capacity limits, calendars, MOQs, service promises, and how often the decision is revisited; and (3) the valuation—a money-denominated score that prices upside and downside under uncertainty. Once these are explicit, the task ceases to be “plan the supply chain”; it becomes “choose among admissible options to maximize risk-adjusted return”.
This decomposition is not a rhetorical device; it is the engineering move that makes unattended software possible. Replenishing a SKU at one location over a given horizon, selecting today’s price for a slow mover, assigning cartons to a truck stop, accepting or deferring a repair on a rotable—each is a specialized, auditable resource-allocation decision. Loosely coupled problems can be handled greedily at high frequency; tightly coupled ones can be grouped and optimized with heavier machinery. Boundaries remain revisable—when economics shift, we widen the option set or adjust the valuation—but while they hold, the decision can be automated, inspected, and improved by a system of intelligence.
The history of artificial intelligence is a game of moving goalposts. Researchers pick a hard problem, label its solution “real AI”, then—once solved—rename it and move on to the next challenge. Each triumph is soon judged mundane, and the cycle begins again.
This pattern explains why mature solutions often look “unintelligent”: they become routine and mechanical. Intelligence should be judged by the quality of choices, not by how arcane the machinery appears. Sophisticated models can masquerade as “smarter”, yet they still address a bounded task; true general intelligence questions must be confronted separately.
Counterintuitively, the first workable solution is rarely the simplest; achieving simplicity is a separate, often harder, task. Progress does not run from simple to sophisticated but from messy prototypes to streamlined cores. Maxwell’s electromagnetic theory began with more than twenty equations; Heaviside later condensed them into the four we use today. In the same way, today’s convoluted answers are tomorrow’s elegant principles.
The same cycle fuels AI booms and busts: people confuse solving a tough niche problem with cracking general intelligence, and investors ride the resulting hype. General intelligence is hard, but countless specialized problems must be harder. Indeed, if we ever build a general intelligence, every remaining unsolved problem will, by definition, be tougher; otherwise they would have fallen first.
The highs and lows of the AI label should not eclipse computer science achievements since World War II—many central to supply chain. The field is now so broad that any overview is necessarily partial, but the following pages single out the topics most relevant to supply chain work.
Algorithms
Algorithms are precise procedural recipes that terminate and return a correct answer for a well-specified input. Formalized in the 1930s (Church, Turing, Kolmogorov), algorithms became the backbone of 20^th^ century computer science. Many early “AI” successes of the 1950s—topological sorting, graph routing, linear-equation solvers—are now simply known as algorithms.
Formally, any program that halts on all admissible inputs could be called an “algorithm”. In practice, computer scientists use the term more narrowly. An algorithm is an atomic, reusable building block whose problem statement is independent of any dataset, interface, or server, accompanied by a proof that, for every valid input, it returns the specified output and by an analysis of its time and space requirements as input size grows.
An “atomic” task cannot be decomposed into smaller, independent, purposeful subtasks. By contrast, rendering a web page involves hundreds of steps—from tokenizing HTML to adjusting character kerning—and thus is not atomic.
An algorithm comes with a proof of correctness; absent such a proof, it is merely a heuristic. The combined Fermat–Fibonacci tests, for instance, detect primes reliably yet still lack a proof of correctness, so they remain heuristics.
An algorithm’s value ultimately lies in its performance. While brute-force methods exist for most tasks, a good algorithm reaches the same result with far less computation and memory. On occasion one can even prove optimality, though the subtleties of complexity theory lie beyond this chapter.
Algorithms (and the data structures that support them) are the bedrock of software; every method discussed later relies on them. Accordingly, they are an auxiliary science for supply chain—though well-designed tools can shield practitioners from their details, as we shall see.
Readers seeking depth should consult Cormen et al.’s classic Introduction to Algorithms (3rd ed., 2009). The annex provides a worked example—the topological sort algorithm.
The CRUD pattern—introduced in Chapter Information—owes its success to the fact that it lets teams ship business software without deep algorithmic chops. In enterprise IT, this became a career track: many capable developers spend decades building forms and reports while never learning to design and analyze algorithms. That gap is harmless inside systems of records, where constant-time CRUD dominates and the database shoulders the heavy lifting; it becomes crippling the moment we leave ledgers and attempt optimization. It also explains CRUD’s cultural dominance and, symmetrically, why vendors brilliant with records so often stumble at systems of intelligence: the latter demand algorithmic modeling, complexity control, and performance engineering—skills CRUD habits neither teach nor select for.
Operations Research
Born in World War II, operations research tackled wartime resource-allocation80 questions—where to place radar, how to route convoys, how to assign scarce aircraft—and, after 1945, business ones. It spawned many algorithms—most famously Dantzig’s 1947 simplex method—but, unlike computer science, it cared less for atomic problems and more for end-to-end applications. In 1947, Scottish radio engineer Robert Watson-Watt defined operations research:
To examine quantitatively whether the user organization is getting from the operation of its equipment the best attainable contribution to its overall objective.
Watson-Watt’s wording echoes this book’s opening definition of supply chain; operations research is its forerunner. Between the 1950s and 1970s, the field produced simulations, econometrics, stochastic models, and mathematical optimization. Many of these strands later grew into disciplines of their own, drifting away from the operations research umbrella.
The most notable contributions of operations research—at least among those not reclassified into their own fields—are methods for transportation optimization, linear and quadratic programming, integer programming, and similar tasks. All these tasks share a common structure: decision variables that must be chosen to maximize an objective function that represents a real-world goal. This framing still underpins most business decision-making, and this book’s approach.
By the 1970s, operations research was falling short of its ambitions. As mentioned earlier, American operations research pioneer Russell Ackoff captured this sentiment in his 1979 paper The Future of Operational Research is Past. Most resource-allocation decisions were still manual in 1979—and most remain so today. Ackoff listed several root causes:
-
Poor systems thinking: experts relied on simplistic, first-order causality.
-
“Pseudo-optimal” results driven by contrived objective functions.
-
Fragile answers that assume a perfectly known, unchanging context.
-
Sanitizing real-world “messes” into toy problems, leaving practice untouched.
Those critiques still apply. Much of the supply chain literature repeats the same mistakes—one reason so little of it survives contact with reality81.
During the 1980s, operations research morphed into “OR” (the acronym), specializing in software “solvers” for optimization models written in domain-specific languages. OR thus settled as a subfield of computer science.
Commercial solvers form a healthy software niche and handle many design tasks well. Nocedal and Wright’s Numerical Optimization (2006) is still the standard survey, even for learning-based variants. Beyond vehicle-routing variants, solvers rarely appear in supply chain because three lingering limitations curb their value.
First, they focus exclusively on deterministic rather than stochastic optimization, making it impossible to handle uncertainty—a constant feature of supply chains.
Second, they offer either scalability or expressiveness—but not both. Solvers scale for supply chain purposes only in linear or quadratic regimes, which are too simple to reflect real-world supply chain situations.
Third, solvers prize mathematical purity over practical expressiveness82. Linear and quadratic programming can indeed certify optimality, yet they describe the real world so poorly that the “optimal” answer is invariably useless.
Setting purity aside, solvers still scale poorly. Faced with a concrete instance—say, vehicle routing—companies often build ad hoc code instead of using off-the-shelf tools. The issue is acute in supply chains, where tens of millions of decision variables (several per SKU) are common.
Thus, the legacy of operations research lies largely in the fields that originally emerged from it. Present-day OR has little to do with the goals operations research set out to achieve. Furthermore, while OR’s favored techniques have merits and notable achievements, they are unlikely to circle back into supply chain.
Statistical learning
The notion of algorithmically inferring a process from recorded observations appeared in the 1950s. It emerged alongside the earliest writings on artificial intelligence. Because individual observations can encode partial solutions, learning has long overlapped with optimization83. Early work on artificial neural networks was once published under the operations-research banner; today it is firmly classed as “learning”.
Until the early 1990s, interest in learning techniques was limited for two main reasons. First, computers lacked the power to achieve anything substantial by learning. Consequently, non-learning approaches—such as expert systems84—were favored. For example, Deep Blue—the first chess-playing program to defeat the reigning world champion, Garry Kasparov, in 1997—was engineered as an expert system. There was also a scarcity of public training datasets, since collecting and distributing large datasets remained prohibitively expensive until the late 1990s. The MNIST dataset contains 60,000 training images and 10,000 test images, each a 28×28 grayscale rendering of a handwritten digit. It provided a shared baseline that fostered a generation of researchers and standardized validation and benchmarking.
Second, high-dimensional learning remained baffling until the late 1990s. Classical statistics assumes the number of variables is small relative to the number of observations. With 100 observations, a two-variable regression can be trusted; with ten variables, spurious patterns proliferate; with 10,000 variables, standard methods collapse. Paradoxically, more information produced less insight.
The curse of dimensionality found its first convincing answer in Vapnik–Chervonenkis (VC) theory, developed from the 1960s through the 1990s85. The theory led to support-vector machines—first implemented in 199586—and provided a mathematical framework to curb overfitting. Support vector machines soon demonstrated their value on real-world datasets.
VC theory reframed the aim of statistical modeling as maximizing generalization—performance on unseen data. Though we cannot measure accuracy on data we do not have, VC theory proves an upper bound on the true error: empirical error plus a model-complexity term determined by the model’s design. This formalization refines the classical bias–variance tradeoff: low-capacity models exhibit high bias, high-capacity models high variance. Figure \ref{fig:under-over-fitting} illustrates underfitting, a correct fit, and overfitting. This theory paved the way for statistical models that can deliver reliable results even when the number of input dimensions vastly exceeds the number of observations.
Operationally, support vector machines were rapidly superseded by alternative techniques—(gradient) boosting (1999)87 and bagging (2001)88—whose mathematical analyses likewise showed they were well-behaved with respect to overfitting. Together, these ideas gave rise to gradient-boosting machines, which remained state-of-the-art across many tasks for roughly a decade. Today, gradient boosting machines remain popular, particularly when observations are scarce.
These developments had an outsized impact: the built-in capacity control of boosting, bagging, and SVMs let theoreticians prove tight bounds on overfitting—assurances rival approaches lacked. Throughout the late 1990s and much of the 2000s, this statistical seal of approval led researchers to sideline nearly every competing approach. Gradually, however, it became clear that no amount of theoretical polish or engineering finesse could push these methods beyond certain hard limits. The technology had effectively stalled. French-American computer scientist Yann LeCun captured the mood by dismissing them as little more than “glorified template matching”.
Despite these limits, statistical-learning ideas remain a cornerstone of modern statistics and a vital auxiliary science for supply chain. For an accessible, figure-rich survey, see the textbook The Elements of Statistical Learning (2009) by Hastie et al. Boosting- and bagging-based models still excel when data are scarce but high-dimensional, and they require far fewer resources than most alternatives—although they do not provide a route to general intelligence.
Deep learning
In the 2010s, deep learning became the dominant machine-learning paradigm. Deep learning descends from artificial neural networks, but it crystallized only after the field abandoned biological metaphors and embraced silicon hardware. Remarkably, its core ideas are simpler than those of its bio-inspired predecessors.
Conceptually, a deep-learning model is a fixed circuit of floating-point operations parameterized by floating-point numbers. Most layers perform matrix multiplications punctuated by rectifiers. Rectifiers break linearity: stacking linear maps remains linear, so a non-linear activation is required. Additional components—convolutions, pooling, or softmax layers—allow the model to handle diverse input modalities and output formats.
Consider again handwritten-digit recognition on the MNIST dataset. If you flatten each 28×28 image into 784 numbers and pass them through three weight matrices with rectifiers, you obtain a 234,752-parameter multilayer perceptron—tiny by today’s standards (see Figure \ref{fig:perceptron}). Running it (performing inference) requires only three matrix multiplications and two rectifier activations (see the annex for details).
So far we have specified the architecture and noted that inference reduces to three matrix multiplications and two rectifiers; what remains is to obtain the numbers inside those matrices. In practice we want a well-trained network—one that, for each input image, assigns its top score to the digit actually present. Training is therefore a search over the parameter space for values that minimize a chosen classification loss. Deep learning tackles this search through three coordinated steps: Initialization—choose sensible random starting weights so activations stay numerically well-behaved; Gradient—compute how an infinitesimal change in each weight would change the loss; and Descent—use those gradients to nudge the weights step by step toward lower loss. Together these steps iteratively refine the network’s weights until performance stabilizes.
Initialization – Weights begin as Gaussian noise and are immediately rescaled so that, layer after layer, activations retain roughly zero mean and unit variance. This simple precaution prevents the product of many linear operations from blowing up (overflow) or shrinking to numerical dust (underflow). When no closed-form rule applies, a quick trial pass through the network rescales empirically to preserve that invariant.
Gradient – The gradient tells each parameter which way—and how far—to move to lower the chosen loss. Modern frameworks obtain it by automatic differentiation, which threads derivatives through the computation graph at roughly the cost of one extra forward pass, but with higher memory since intermediate activations must be cached. Earlier deep learning toolkits used hand-coded gradient back-propagation; that the community took decades to adopt automatic differentiation — dating from the 1960s — is a reminder that progress is rarely linear.
Descent – Gradients become updates through stochastic gradient descent (SGD), a 1950s-era workhorse that still dominates deep learning. SGD samples one observation—or, for better hardware utilization, a small set of observations referred to as a mini-batch—nudges the weights in the gradient direction, then repeats. Mini-batching trades extra arithmetic for sharply lower wall-clock time by exploiting parallel cores. Variants with momentum, adaptive learning rates, or decoupled weight decay refine the basic update.
Deep learning overturned several statistical-learning dogmas and still shapes what can—and cannot—be inferred from company data.
The double descent
Deep-learning models are hyper-parametric: they usually contain more parameters than training examples. The 234,752-parameter perceptron above dwarfs MNIST’s 60,000 images. Statistical theory predicts severe overfitting, yet practice says otherwise.
Discovered in the late 2010s, the double descent curve helped clarify why many statisticians initially doubted deep-learning claims. Trained to expect the classic U-shaped bias-variance trade-off, they were puzzled when larger networks sometimes outperformed smaller ones instead of overfitting. Double descent shows why: as model capacity grows, test error first drops (good generalization), then climbs (over-fit), but beyond a second threshold drops again as very large, overparameterized models settle into new low-error regions (see Figure \ref{fig:double-descent}).
On the left of the double descent curve—the classic regime—the network has too few parameters, so it underfits. As we add weights, the test loss falls until it reaches a first minimum at the “sweet spot”. Push past that point and the model swings into classic overfitting: extra capacity now memorizes noise faster than it learns signal. Curiously, in this zone the training error continues to slide smoothly even as the test error rises—a behavior that long puzzled statisticians who expected the two curves to diverge much earlier.
Move to the right—the deep regime—and the pattern flips. Once the parameter count crosses a second, much higher threshold, test loss starts dropping again: larger really is better, as deep-learning practitioners had claimed since the early 2010s. Yet each extra slice of accuracy costs an outsized chunk of compute and energy, so returns diminish rapidly. State-of-the-art networks now operate well beyond that threshold, trading massive hardware budgets for the last fractions of a percentage point in performance.
For supply chains, the message is mixed: giant models win accuracy but hurt maintainability, while smaller ones stay agile yet fall back into the classic under/overfit dilemma. Hybrid designs that mix the two are frequently the superior option.
Mechanical sympathy
Artificial neural networks (ANNs) have a clear weakness: they are grossly inefficient in their use of computing resources. Silicon chips resemble biological neurons so little that even a crude digital neuron squanders cycles. Deep learning broke free of this handicap by adopting what race engineers call mechanical sympathy. In early Formula 1, the fastest laps were set not by drivers who ran at the limit, but by those who sensed how much stress the engine, gearbox, and brakes could tolerate and modulated their pace so the car finished intact at the checkered flag. Translated to code, mechanical sympathy means arranging algorithms and data to move with the grain of caches, vector units, and memory buses instead of colliding with them. Once this mindset took hold, researchers shed nostalgic neuron metaphors and kept only what maximized raw computational throughput.
For example, early ANNs used the sigmoid activation function, chosen to echo biology, but its exponential is slow in hardware. The simpler rectifier (ReLU) proved faster and quickly took over.
In the early 2010s, treating cache lines, vector widths, and memory buses as first-class design constraints felt almost heretical. The algorithmic, operations research, and statistical learning camps had risen on hardware-agnostic theory: prove a better asymptotic complexity class and let Moore’s law absorb pesky constant factors. Tuning for those constants—say, shaving cycles with a rearranged matrix or a SIMD instruction—was seen as busywork, soon nullified by the next processor generation. Deep-learning researchers upended that view, showing that painstaking attention to hardware minutiae could yield speedups unattainable by elegance in algorithmic complexity alone.
The early quest for ever-larger networks was fueled by the biological metaphor: with roughly 100 billion neurons, the human brain became an informal size benchmark for ANNs. Experiments soon made the metaphor unnecessary—performance rose simply because the models grew. Meanwhile, the data landscape flipped. In the 1990s, public datasets were scarce; by the 2010s, they were expanding faster than networks could absorb them. Training on every available sample became impractical, and vast troves of information remained untapped.
Hence the arms race: ever-larger models fed by ever-larger datasets. Toronto’s 2012 ImageNet network had 60 million parameters and ran on GPUs; a decade later, trillion-parameter transformers sit on custom silicon. The scale boom also spawned heavyweight tooling. Google’s TensorFlow, released in 2015, ballooned past four million lines of code—an industrial effort unthinkable in the statistical learning era of 10,000-line libraries.
The race for ever-larger models also pushed designers toward ever-smaller floating-point formats. Because transistor count, and therefore energy cost, grows faster than bit width, shaving precision delivers outsized speedups: on current GPUs, arithmetic throughput can quadruple when you drop from 32-bit to 16-bit. Early 2010s deep-learning systems used 32-bit floats exclusively—a safe default that still suffices for most supply-chain calculations, provided pathological edge cases are avoided (some tasks still benefit from 64-bit). By the early 2020s, however, training in 16-bit and serving in 8-bit had become routine, courtesy of a friendly co-evolution of GPU and CPU hardware. Research is now probing 1-bit—and even sub-1-bit—weights, showing that the lower bound on usable precision has yet to be reached.
Mechanical sympathy matters in supply-chain practice. Swiss computer scientist Niklaus Wirth warned in the 1990s that software often slows faster than hardware accelerates—a pattern now called Wirth’s law. Because of a general lack of interest in hardware efficiency, many business systems run no faster today than in 2000, despite far better chips. That slowdown is Wirth’s law at work. Deep-learning tricks cannot be reused wholesale, yet they reveal how mindful coding unlocks huge gains. Supply-chain professionals therefore need at least a working grasp of mechanical sympathy.
Nonconvex optimization
In the early 2000s, mathematical optimization still followed the original operations-research playbook. Three tenets defined it. Statistical-learning practitioners—support-vector-machine devotees included—accepted it wholesale.
First, early optimization theory treated local minima as a practical brick wall: in nonconvex landscapes one might, in principle, escape them, yet in real projects one almost never did. That assumption spawned an entire branch—convex optimization—whose express purpose is to sidestep that obstacle. Because a convex function has no local dips—only a single global minimum—any well-behaved descent algorithm converges to it. For a thorough treatment of the proofs, algorithms, and applications that rest on this property, see Boyd and Vandenberghe’s Convex Optimization (2004).
Second, for nonconvex problems, any respectable method was expected to provide a bound on the distance of its solution from the global optimum. One does so by furnishing an upper bound on the gap between the best-known solution and the global minimum. Techniques such as branch‑and‑bound and branch‑and‑cut offer such guarantees. However, unlike convex optimization—which benefits from highly scalable techniques—these general methods scale poorly. As mentioned earlier, the poor scalability of such solvers has long been a chief reason for their limited adoption in supply-chain circles.
The deep-learning community jettisoned that classical view. It embraced a blunt tool—stochastic gradient descent—and applied it to nonconvex landscapes, sidestepping fears of local minima. This approach has proved wildly successful, enabling models with trillions of parameters—a scale inconceivable in the 2000s.
A new outlook emerged: the obstacle was no longer local minima but broad, flat plateaus. The prevailing view holds that local minima become exceedingly rare as dimensionality increases. Thus, intuition from low-dimensional examples fails in high dimensions89. Instead, vast low-gradient regions (“plateaus”) pose the main challenge for descent methods.
Third, for (hyper)parametric models, optimality proofs become meaningless in the double-descent regime. The training loss can plunge and hover near zero, yet those tiny residuals say little about how the model will generalize. Worse, many loss functions favored in deep learning—such as cross-entropy—do not even have true minima90.
While the deep-learning community had reasons of its own to reject the classic optimization stance inherited from operations research, this rejection matters for supply chain. It shows that the classical paradigm suffers from counterintuitive limitations that escaped notice for roughly six decades (1950s–2000s). In what follows, we will see that supply chain has further grounds to reject this classic stance and that alternative, far more valuable paradigms exist.
The bag of tricks
Every programming paradigm offers its own mental model of software work, with core concepts, practices, and tools; most are opinionated about data representation and control of program behavior.
Deep learning was the first91 broadly adopted programmatic learning paradigm; in the early 2010s, it puzzled researchers by defying prevailing expectations in machine learning.
Until then, papers and software focused on the finished product: the model ready to be trained on a fresh dataset. What set deep learning apart was its focus on architectures, numerical tricks, and guiding principles for crafting models. While toolkits for statistical learning (such as scikit-learn92) featured dozens of models, those for deep learning shipped with none. The field spent much of the 2010s absorbing that shift. The old approach—identifying a novel model, running a benchmark, and publishing results—became obsolete. Novelty no longer lay in the model itself or its statistical characterization. Even first-year computer students could cobble together a “unique” deep learning network by cherry-picking layers from the available tensor operations and benchmark it on one of many public datasets. For a brief period, deep learning conferences and journals were flooded with papers that were little more than such compositions; despite their “uniqueness”, such contributions were often inconsequential.
Deep-learning architectures serve as the modern analogue of the “models” once championed by operations research, yet they leave significant latitude for the data scientist in finalizing the network to be trained and deployed.
The most valuable contributions came to be known as the “bag of tricks” of deep learning: small, modular tweaks that improve parts of a network rather than redefining the whole. Weight initialization strategies, moment estimators, learning rate scheduling, dropout, early stopping, residual connections, dense connections, attention mechanisms, transformer architectures, batch normalization, gradient accumulation, hyperparameter optimization, data augmentation, positional encoding, pruning, quantization, knowledge distillation, low-rank factorization, among others, form only a small subset of this growing “bag of tricks”. However, these contributions—at least the significant ones—proved highly accretive. One technique seldom precludes another unless they belong to the same class.
Since the 1960s, many paradigms—functional, object-oriented, distributed, reactive, and more—have tackled general software concerns. Specialized paradigms have also emerged, such as relational algebra (e.g., SQL for relational databases). While debates continue over which paradigm is “best”, no serious software project is undertaken without embracing one or more paradigms. Merely choosing a programming language already constitutes a paradigmatic choice.
The key takeaway for supply chains: choosing the right paradigm matters even outside pure software work. Deep learning, as a programming paradigm, has redefined the state of the art in machine learning. This raises the question of whether other domains, such as supply chain, could benefit from similarly dedicated paradigms; as we shall see, dedicated programming paradigms are of primary importance for supply chain.
Generative Era
In 2022, a handful of breakthrough releases—soon grouped under the banner “Generative AI”—vaulted from niche research to global headlines. ChatGPT, the conversational assistant unveiled by OpenAI in November 2022, and Stable Diffusion, the text-to-image model released in August 2022 by researchers at LMU Munich and Heidelberg University, became instant landmarks in text completion and image generation, respectively. Their apparent overnight success masked years of steady, open-ended progress by the wider deep-learning community; what changed in 2022 was not the underlying science, but the packaging. For the first time, non-specialists could summon large, pre-trained models through a simple web interface and obtain striking results in seconds—a watershed that turned an academic curiosity into a mainstream product category. Although these systems are squarely built on deep-learning foundations, they also challenge some of the paradigm’s early assumptions—for example, about data quality and loss-function design. In that sense, the “generative era” represents a genuine second generation of deep-learning practice rather than a mere extension of the first.
First, vendors began shipping pre-trained models by default—a radical, albeit unexpected, advance. With pre-trained models, end-users need no longer concern themselves with generating models (the learning phase) and can focus on using them (the inference phase). This greatly eased adoption. It also paved the way for hybrid learning processes, in which a deep learning network could be assembled from pre-trained components and supplemented with a trainable section. Although staged learning processes predate deep learning, the generative paradigm added a new dimension and prompted the machine learning community to adopt further open-source practices, sharing models in addition to code.
Second, the hallmark of the generative era is an almost insatiable appetite for web-scale data. For decades the adage was “garbage in, garbage out” (GIGO): low-grade inputs doom a model’s outputs. Yet today’s text- and image-generators train on billions of lightly filtered pages and pictures scraped straight from the open web, and still return prose and visuals that eclipse the average quality of that raw material. Their results are hardly flawless, but they are unmistakably better than the median of their sources—an outcome that overturns the older GIGO credo. Harnessing such oceans of data, however, is only possible when the software is written with deep mechanical sympathy, squeezing every cycle and memory bank the hardware can offer.
Third, ever-larger models display emergent skills for which their loss functions were never designed. For example, once large language models (LLMs) reach billions of parameters and are trained on billions of tokens, they start exhibiting translation capabilities. Yet the loss function reflects only next-token prediction; it does not account for translation. Similarly, text-to-image models exhibit emergent capabilities such as style transfer that go beyond their training data.
Those abilities upend an assumption borrowed from operations research. Until then, it was generally assumed that the best way to solve an optimization problem was to steer the process by the very metric used to evaluate the final solution. Proxy loss functions—approximations of the true loss—were acceptable only when they offered significant computational benefits—a trade-off between solution quality and resources. However, the emergent capabilities of generative models have repeatedly surpassed prior state-of-the-art methods driven by more direct loss functions.
In the generative era, the relationship between the score for a given task and the loss function used to drive stochastic gradient descent has become tenuous. Self-supervised learning has supplanted supervised learning, which had dominated until then. The intuition behind self-supervised learning is that the quantity of unlabeled data93 typically dwarfs the amount of labeled data. Thus, if learning is decoupled from the narrow task of predicting “labels” and instead focuses on intrinsic properties of the data, it can better capture underlying structure.
While generative models have redefined the state of the art in machine learning, leveraging them for supply chain purposes is not as self-evident as many software vendors claim. Relational records constitute the bulk of data relevant to supply chain, yet neither text completion, text-to-image, nor image-to-text models are well suited for processing such data. Two propositions commonly put forward by vendors in the early 2020s warrant refutation.
First, large language models fit text, not numerical time-series forecasting. Although there are proposals94 to enhance the arithmetic capabilities of such models, these developments are unlikely to render language models competitive with models specifically engineered for numerical processing. At best, large language models may become an extremely convoluted way to achieve what simple parametric models can accomplish with a tiny fraction of the computing resources.
Second, chat-style interfaces usually hurt productivity. Conversing with a chatbot is significantly slower than traditional navigation. Any task intended to be repeated routinely deserves a “traditional” user experience rather than a cumbersome conversational interface. In an enterprise environment, employees are expected to be trained and proficient with the system—not to “muddle through” it. The enterprise perspective differs fundamentally from the end-consumer perspective, where users are assumed to be less knowledgeable regardless of the task.
In supply chain, large language models shine at drafting numerical recipes and smoothing the informal work that surrounds decisions; we will revisit these contributions later.
The Future
Every supply chain decision carries an implicit claim: “we expect the company’s future conditions to make today’s allocation of resources profitable.” The prediction may remain unsaid, but it cannot be evaded. A firm may ignore the forecasting challenge; it cannot ignore the consequences.
Thus, the future—and, more generally, time—matters greatly to supply chain. Yet time is deceptively difficult—not because it is too abstract, but because it appears self-evident and resists deliberate thought. The future is akin to the air we breathe: ubiquitous and yet invisible.95 With such topics, one is tempted to take preconceptions for granted and skip the intellectual work altogether. Chemistry taught us that millennia-old preconceptions about air were wrong96. Just so, in supply chain, unexamined beliefs about tomorrow—often a tacit extrapolation of yesterday—quietly shape prices, capacities, and inventories. When they are wrong, they harden into plans that misallocate scarce resources and propagate failures across the flow.
To anchor the discussion, this chapter connects the simple observation—every decision is a bet on tomorrow—with the instruments that make such bets profitable. It proceeds in three moves.
First, we surface the unspoken visions that steer how firms think about the future. Much twentieth-century practice descends from a teleological vision: settle tomorrow by forecasts and an engineered plan. By contrast stands a rugged vision: treat tomorrow as a field of opportunities under uncertainty and preserve options so the firm can move when the odds turn favorable.
Second, we examine why forecast-driven planning collapses in business even when it looks mathematically tidy: time-series formalism flattens what matters; the future is shaped by decisions not yet taken; small numbers, batching, adversarial behavior, and fat tails shatter the neat symmetry between past and future. The result is brittle orchestration and bureaucratic ritual.
Third, we assemble predictive instruments that fit the rugged vision and the economics of optionality introduced earlier in the book. We replace point time series with probabilistic forecasts (demand, lead times, prices, returns), add high-dimensional forecasts (market baskets, substitution, contagion), and introduce functional forecasts that speak in the native objects of the flow rather than in equispaced time buckets. These tools do not guess “the” future; they price alternatives and make opportunity costs explicit.
Visions
[A vision] is what we sense or feel before we have constructed any systematic reasoning that could be called a theory [..] A vision is our sense of how the world works [..] Visions are all, to some extent, simplistic – though that is a term usually reserved for other people’s visions, not our own. A Conflict of Visions (1987), Thomas Sowell.
Of all the unspoken ideas that drift around us, the most powerful shape our intuition of causality for the objects at hand—here, supply chains. Indeed, this intuition defines how we frame situations, how we see problems—whether we see them at all. Here, vision refers to that intuition of causality.
Vision precedes theory, which precedes models. Visions permeate companies—their cultures, practices, and processes. Visions matter because they serve as a rough compass for thought. Amid the infinite subjects that could claim our attention, visions decide what is worth thinking about.
For example, “Day 1”, as advocated by Jeff Bezos, Amazon’s founder, is one such vision. It emphasizes a perpetual startup mindset—continuous innovation and proactive action—to avoid the stagnation of “Day 2”. It means treating every day as if it is the first, encouraging rapid decisions, embracing experimentation—even at the risk of failure—and favoring long-term opportunities over short-term gains. By adhering to the “Day 1” mindset, Amazon aims to sustain its entrepreneurial spirit and avoid the complacency that plagues larger, more mature organizations.
Yet the opposite vision also has its proponents. As Captain Chesley “Sully” Sullenberger—famed for the Hudson River emergency landing that saved 155 lives—remarked, “For 42 years, I’ve been making small, regular deposits in this bank of experience, education, and training. And on January 15, the balance was sufficient for a very large withdrawal.” This view dominates aviation, combating complacency not through disruptive innovation but through relentless refinement of existing practice. Past failures were paid in blood, and future ones will be as well.
Visions govern how people instinctively approach complex systems—systems beyond what a human mind can readily grasp in full. Consider a retail store struggling to maintain adequate stock, with half its shelves empty. Instinctive diagnoses will vary greatly, depending on one’s vision of how supply chain works.
Take a professor of supply chain analytics. He may instinctively blame inaccuracies in the demand forecasts driving replenishment. Here, the vision places the burden of service quality on a technological solution—on software. This vision extends to his academic community, whose research influences the design and accuracy of such software, thereby reinforcing a technology-centered view.
By contrast, a regional manager in the same chain may blame poor store stewardship. In this vision, people—the store manager and staff—are responsible for running the store efficiently. Responsibility lies first with those closest to the problem. This vision also implicates upper management: they allow an ineffective store manager to persist in his position, further underlining a people-focused view.
These two visions arise from the same empty shelves but assign responsibility—and, consequently, resolution—to entirely different quarters. One turns to a technological solution, the other to a management solution. A third, equally plausible diagnosis lies outside the firm: rampant shoplifting makes displayed inventory vanish. Naturally, whether the store’s problem stems from faulty software, poor management, or an adverse public-order context is another matter entirely. Visions prove nothing; they merely govern our immediate assessment of complex situations.
Visions are consequential. History shows that many businesses came to dominate their markets thanks to the peculiar visions of their founders—“visionaries,” aptly named97. Yet the same visions can, over time, set companies up for failure. Thus, for any company, it is reasonable to identify the dominant visions held by upper management and assess their fitness for the business.
Unlike corporate strategies, visions are rarely discussed, let alone weighed for merits and demerits. Within the same company, people may hold radically different visions. Because visions are rarely spelled out, employees feel that whenever they push, others pull—and vice versa. Yet divergence in vision is seldom named as the cause.
Values
In politics as well as in business, leaders often highlight divergent values to underscore differences with their rivals. The phrase “We do not have the same values” can be heard on all sides. Yet this perspective, while not without merit, tends to obscure the profound influence of visions.
In A Conflict of Visions (1987), Thomas Sowell astutely notes that when people meet differing interpretations of the same facts, they tend to attribute the differences to values. Often the variation in values is far less pronounced than the catchphrase “We do not have the same values” suggests. In the political realm, for instance, one would be hard-pressed to find anyone advocating for poverty, crime, or war. Yet despite shared values against these ills, people’s visions guide them toward starkly different solutions.
The same holds in supply chains. Regardless of field or sector, companies seek profitability, growth, service quality, and waste reduction. Companies that openly oppose such values are rare, if they exist at all. Different visions yield markedly different strategies and practices, all aimed at the same values.
Consider Amazon’s founder, Jeff Bezos, who has often emphasized a relentless focus on the customer. He once said, “The most important single thing is to focus obsessively on the customer. Our goal is to be Earth’s most customer-centric company.” That is a statement of values—and a fairly mundane one. Indeed, executives rarely devalue customers in public. When an executive is caught doing so, he rarely keeps his position afterward. What distinguishes Amazon is not its espoused values—which align with most businesses—but its distinctive vision and culture.
Thus, as we assess how understanding the future shapes supply chain practice, we must remember that men may propose radically different solutions while sharing the same values. More often than not, the divergences are rooted in conflicting visions.
Time
The passage of time shapes our experience and our understanding of the world. It is so concrete and omnipresent that abstracting away the dimension of time is difficult. Yet many—if not most—supply-chain challenges deserve this treatment, if only to surface the underlying visions at work with regard to time. Thus we pose a series of questions about time. At this stage, answering those questions is unimportant and will not be attempted. Instead, we aim to characterize the visions—the intuitions of causality.
Do we see the future as knowable or unknowable?
Forecasting is only as valuable as the portion of tomorrow that can be anticipated in practice. When the future is at least partly knowable, forecasts are powerful instruments; when it is dominated by noise, they are mere ceremony. Planning inherits the same premise: it pays only if some features of the future can be foreseen reliably enough to guide commitments. If outcomes routinely diverge from expectations, planning effort turns to waste.
Do we see the future as fated? Or open to change?
If the future is fixed, a forecast is an early measurement of a state that will occur; accuracy is judged by its match with the realized outcome. If the future is malleable, a forecast acts as a directive shaping behavior; accuracy is judged by how closely execution conforms to the plan.
Do we see uncertainty as a flaw or as risks and opportunities?
Few dispute that the future is uncertain. One view treats that uncertainty as a defect to be minimized—via sharper forecasts, tighter controls, and prompt corrective actions. The other treats it as raw material—to be priced, hedged, and occasionally embraced when odds favor it.
Do we see acting sooner as preferable to deferring until a later, better-informed decision?
Momentous decisions made on the spot strike some as rash and others as decisive. Likewise, inaction reads either as weak-willed hesitation or as strength to wait for better information.
Do we see knowledge as a characteristic of what doesn’t change, or as the specific circumstances of time and place?
Scientific knowledge aims at permanence: theories are cast as timeless claims, and reproducibility assumes a phenomenon that does not change with date or venue. Business knowledge is the opposite. Once a fact is widely known, its private value collapses. What creates an edge is the specific circumstances of time and place—who wants what, at what price, given today’s capacities and constraints. Valuable knowledge is therefore local, timely, and perishable.
Do we see the past as a foundational asset to build upon, or a flawed legacy steering people toward complacency?
Founders may start from a blank page; every successor inherits a status quo. One stance treats that legacy as a compounding asset—skills, supplier ties, data, and routines that, when refined, reduce risk and cost. The other treats it as a liability—sunk costs, ossified habits, and internal constituencies that bias choices toward yesterday’s market.
Do we see investment in people as indefinitely accretive, or is accretion reserved for technologies and techniques?
Some see human capability as compounding. With practice, feedback, and better mental models, planners keep getting faster and more accurate; thus spending on training is a capital-like outlay. Others expect an early plateau: once competence is reached, durable gains come chiefly from improved techniques and software; investment should therefore target tools rather than people. The stance taken strongly governs what counts as an “investment” in the first place.
These questions are, of course, false dichotomies: the truth may lie between the poles, vary with context, or lie elsewhere entirely. Far from mere philosophical musings, these questions show that one’s intuition about time—about the future—is consequential. Diverging visions do not merely produce divergent solutions: people may not even agree on the problems worth solving.
Within supply chain circles, views of time and the future have largely converged into what we call the teleological vision: the belief that tomorrow can be fixed in advance by forecasts and then engineered into existence through compliance with a plan. Over the 20^th^ century, this view became dominant and remains the default in the 21^st^ century. It is seldom named, rarely questioned, and silently assumed by most authors and enterprise software alike. The next section makes this vision explicit and examines its consequences.
The teleological vision
The future remains inaccessible only to those who have not embraced the mathematical instruments that modern science offers. Whatever weaknesses still plague our present-day forecasting models will ultimately be remedied by the unstoppable force of progress. Furthermore, those instruments not only tell us what will come to pass but also why. Thus, the modern leader, unlike his primitive predecessor, leaves nothing to chance. His plans are no longer rooted in the fallible intuition of men; rather, they are engineered to the precision that characterizes this machine age.
These lines capture the essence of the teleological vision98. Over the course of the 20^th^ century, this vision became the dominant view in intellectual circles. By intellectuals we mean people who make a living selling their ideas (writers, historians, academics), as opposed to selling tangible services or products. This definition passes no judgment on intelligence; it simply labels occupations. For example, a journalist is an intellectual, while a neurosurgeon is not. However, the neurosurgeon can be expected to be more educated and more intelligent than the journalist; such is the expectation of every patient placing his life in the surgeon’s hands.
Once a view grips intellectual circles, it often spreads and shapes public opinion99. The teleological vision became dominant—if not exclusive—within supply chain circles, among authors and practitioners alike. Enterprise software vendors entrenched this vision through workflows and technologies that left little or no room for alternative views.
Unfortunately, this vision is as popular as it is flawed. Despite a mountain of contrary evidence, it remains one of the most seductive economic ideas of all time. The prospect of weeding this vision out of supply chain is daunting but necessary. We must first examine where this view came from and how it spawned its supply-chain offshoots.
Historical emergence
Modern quantitative forecasting100—as opposed to the traditional soothsayer, augur, oracle, or prophet—emerged at the beginning of the 20^th^ century in the United States with the first generation of economic forecasters. Demand for forecasts was fueled by the emergence of an investing middle class unique to the United States at the time. At the start of the 20^th^ century, only half a million Americans owned stocks; by the end of the 1920s, that figure rose to 10 million. Those early forecasters were confident that their techniques would ultimately make economic uncertainty a relic of the past:
Week after week, his [Roger Babson] newsletter explained to its readers that economic events were neither random nor uncertain but governed by forces that could be understood through careful study. He began issuing regular forecasts after the Panic of 1907. […] He offered prognostications on everything from prices to employment to sales. —Walter A. Friedman, Fortune Tellers (2013)
Babson’s swagger set the template for economic pundits throughout the century. Babson’s core idea was to adapt Newton’s third law101—also known as the law of action and reaction—to economics. He held that it would ultimately render the course of the economy as predictable as that of celestial bodies. While now widely regarded as naive and misguided, mid-20^th^ century forecasters remained as confident as Babson, shifting their trust from Newtonian mechanics to statistics.
The Soviet Union’s Gosplan, founded in 1921, began issuing annual plans in 1925. The committee drew directly on a teleological vision that the future could be conquered. Yet the committee produced nothing but unrealistic plans that could not be executed. The results proved disastrous. The Soviet economists Nikolai Shmelev and Vladimir Popov write: “The administrative system was distinguished by economic romanticism, profound economic illiteracy, and incredible exaggeration of the real effect that the ‘administrative factor’ had on economic processes and on the motivations of the public.”
Popular culture echoed these beliefs: Isaac Asimov’s science-fiction series Foundation (1951) captures the prevailing views of the era’s intellectuals. The central storyline turns on a fictitious theory, “psychohistory”, blending particle physics and sociology, said to achieve near-perfect predictive power when vast numbers of humans are involved. However, Asimov concedes that some individuals can counter psychohistory’s predictive power. It takes, however, what amounts to a demigod—a human blessed with immense psychic powers—to do so.
Most twentieth-century intellectuals remained supremely confident in the ultimate predictability of human affairs, even as economic forecasts kept failing. While Gosplan’s failure can be attributed, in part, to the general flaws of socialism, market-driven American pioneers such as Moody’s and Standard & Poor’s also suffered setbacks. Both began by selling investment publications—manuals, company digests, and market letters that included explicit forecasts—rather than merely tabulating past accounts. By mid-century, both firms had pivoted to credit rating—a “safer” business that spared them the embarrassment of inevitably faulty forecasts.
Forecasting setbacks did not slow operations research: as the field rose in the 1950s and 1960s, it embraced the teleological view. The future was taken as a known input to the model, either under stationary conditions or via a time-series forecast. By the end of the 1970s, time series had become the cornerstone of quantitative supply chain theories. The prominence of time-series forecasts increased further with the rise of S&OP (Sales and Operations Planning) at the end of the 20^th^ century. Their status and importance remained essentially unchallenged until the early 2010s.
Time-series paradigm
A time series is a sequence of values of a quantity observed at successive times, typically at equal intervals102 between them. Extending the series beyond today yields a forecast—specifically, a time-series forecast. While time series can represent anything—from sunspot counts to a country’s population—most supply chain authors and software use them primarily to represent demand, measured in units (eaches) or in monetary terms.
A time series is an unbounded sequence of numbers laid out on a ruler, with “0” marking today. Everything to the left has already happened; everything to the right has yet to be inferred. In practice, we fill those slots with forecasts, treating time in separate, evenly spaced steps (days, weeks, and so on).
This formalism clarifies a few assumptions that underpin the time-series paradigm. First, time series have no beginning and no end, like a geometric line. Of course, any visual representation is truncated to a finite segment for practical reasons. However, this fragment should not be confused with the whole.
Second, time series are structurally symmetrical with regard to their time dimension. By “structural” we do not mean past and future values coincide, only that they are of the same nature. They are undifferentiated save for the cutoff at time zero, a matter of convention.
Third, time series adopt a discrete time axis. This choice drastically simplifies both the display of time series and the models built atop them, whether analytical or predictive.
These three assumptions match the prevailing view in the natural sciences. A meteorologist studying temperatures at a given location adopts the same paradigm. Though records begin at some date, the meteorologist does not doubt temperatures could have been recorded earlier. Conversely, measurements can continue indefinitely. Although temperature evolves continuously, records are deliberately discretized—for example, one daily measurement at a fixed hour—to ease analysis. Finally, the physical laws are nearly perfectly time-symmetric103 with respect to time.
It is hard to overstate the time-series paradigm’s dominance in supply-chain circles at the beginning of the 21^st^ century. Nearly all forecasting literature revolves around demand time-series forecasting. Few authors acknowledge alternative forecasting targets104—beyond demand—and fewer still acknowledge alternative forecasting structures—beyond time series. Forecasting has effectively become synonymous with “demand time-series forecasting”.
Similarly, most “applied optimization”—the heir to operations research—takes time series as input whenever possible. Again, few acknowledge alternative perspectives.
On the software side, most reporting systems rely on OLAP (online analytical processing) cubes that, by design, embrace the time-series paradigm. Among the cube’s dimensions, one is invariably time. Crucially, such cubes are incompatible with alternative—more expressive—structures beyond time series.
On the management front, the widespread practice of S&OP (Sales and Operations Planning) puts time series squarely at the center. Cross-functional teams convene to produce a single deliverable: the strategic plan. In practice, that plan is a bundle of time series—usually the firm’s projected sales—arrayed across future time buckets.
Planning through time-series
In business, a time-series forecast is not a neutral projection as in the natural sciences; it doubles as a commitment. Once a forecast is sanctioned, it becomes the baseline that sets budgets, reserves capacity, and schedules replenishments. The organization then works to make those numbers hold; the forecast becomes less a guess about tomorrow than an instruction to shape it.
Once forecasts are set—future demand treated as known—all that remains is to orchestrate resources to meet it promptly at minimal cost. In the teleological view, planning means fixing future demand and assigning the matching resources.
Indeed, if the future is treated as settled—one known daily demand for every SKU at every location—then the planning problem collapses into arithmetic: the material decisions (what to buy, make, or move; where; when; and how much) are mechanically inferred from the forecast itself105. Execution then reduces to gauging compliance with “the plan”. For example, if tomorrow’s e-commerce forecast for a given SKU at the fulfillment center is exactly 2 units, stocking fewer is a planned stockout; stocking more is, by definition, avoidable overstock that a more precisely timed replenishment “should” have prevented.
Once the forecast is sanctified as “the plan,” the firm’s agency migrates to the ritual of fixing that number. The forecast sits on the throne; everything else is compliance. Hence the turf wars around the baseline—not a dispute about statistical merit, but a budget negotiation in disguise. To forecasters, the wrangling looks perverse; to managers, “accuracy” is incidental. The real prize is a larger share of capital, capacity, and attention for one’s own fiefdom.
Orchestration then runs into a structural mismatch: forecasts are cast in one geometry, resources in another. The baseline may be weekly by SKU, while trucks depart daily and never on Sundays; production is batched around changeovers; purchase orders obey MOQs and price breaks. A small army of planners is therefore tasked with back-scheduling every detail—shifting quantities across days, squeezing orders before cutoffs, padding here, trimming there—so the ledger appears aligned with the baseline. To an outsider this bustle looks like “making decisions”. In truth, discretion is minimal: choices are bounded by compliance with the plan, and the remaining degrees of freedom are clerical rather than economic.
This view also explains why pricing is almost always ignored by supply chain authors, practitioners, and software vendors. Indeed, because prices affect demand, they belong within supply chain—as we maintain throughout this book—yet the classic stance, following the teleological view, ignores them altogether. Prices are simply expected to be settled prior and independently of the planning phase. To the traditional supply-and-demand planner, prices are a given—neither to be assessed nor challenged.
This view of planning—i.e., a forecast steering the allocation of resources—is so dominant in present-day supply chain circles that few even acknowledge alternatives. Disagreements are narrowed to the choice of forecasting methods, inventory models, and orchestration workflows. Yet, despite its popularity, teleological planning has demerits and deserves scrutiny.
Limits of planning
So let us call here the teleological fallacy the illusion that you know exactly where you are going, and that you knew exactly where you were going in the past, and that others have succeeded in the past by knowing where they were going. Antifragile: Things that Gain from Disorder (2012) by Nassim Nicholas Taleb
Planning, as it emerges through the teleological vision, is deeply flawed. Still, despite a century of contrary evidence—most notoriously the fiascos of the Soviet Gosplan—it remains the central doctrine of large organizations, public and private. Few ideas have seduced the intellectual circles of the 20^th^ century so thoroughly while delivering so little. Before examining why this fascination endures, we first clarify planning’s technical shortcomings, beginning with the inadequacies of time series.
As a mathematical structure, time series are ill-suited to reflect business situations. It is tempting—but profoundly misguided—to recycle time series as used in the natural sciences. The crux of the issue is that, in business settings, time series categorically ignore entire facets of the phenomenon under scrutiny.
Consider the meteorologist taking one temperature measurement every day at the same place—the time series of interest. The meteorologist knows that the measurement itself could be made more precise with slightly better equipment. He knows the measurement could be taken twice, a minute apart, or 100 meters apart. None of these improvements, however, is expected to change the picture radically. It is always possible to make measurements more precise or more granular—but with diminishing returns. The time series is an approximation, yet the meteorologist controls its nature and extent. The meteorologist is never “seconds away” or “meters away” from a radical breakthrough that would upend his understanding of the weather.
Business doesn’t play by those rules. Seemingly minor measurement changes can radically alter the assessment of the situation. Consider a company that has sold the same product for years. Sales are steady—about 1,000 units per week—with little fluctuation. What are the odds that the demand for this product will drop to zero next week and remain at zero forever? To answer this question, let’s consider two situations, both exhibiting identical historical time series of weekly sales.
Consider two demand structures producing the same weekly total of about 1,000 units: Case A—1,000 independent customers, each buying roughly one unit per week; Case B—a single customer buying roughly 1,000 units per week. The historical time series—weekly totals over years—looks identical in both cases, yet the tail risks diverge. In Case A, taking next week’s sales to zero would require losing essentially all customers, a process that—short of catastrophe—unfolds over many weeks or months. In Case B, one decision by one buyer suffices; the hazard of complete collapse is the churn probability of that single account.
Consider the inventory implications. In Case A, declines typically appear gradually, allowing stocks to be trimmed as churn reveals itself; write-offs tend to remain contained. In Case B, a large write-off is a matter of when, not if: once the sole buyer defects, the stock accumulated for them becomes overhang. Thus, two situations with identical weekly sales time series entail opposite risk profiles—and call for different decisions.
Consider a second setting: a store sells an average of 10 units per week of a given SKU. The store is open seven days and replenished daily. What on-hand level delivers reliable service? The point forecast for daily demand is $$\approx 1.4$$ units, yet two demand structures sharing the same time series imply opposite stocking needs.
Variant A — many small baskets. Ten distinct customers each buy one unit during the week. With daily replenishment, keeping roughly 5–6 units on hand suffices for most days: one day of cycle stock plus a modest buffer covers ordinary noise.
Variant B — few large baskets. One customer per week buys 10 units at once. Holding fewer than 10 units guarantees a stockout for that visit; to achieve decent service, you need at least 10, and because two such customers may arrive on the same day, 20—possibly 30—is the relevant range.
The time series is identical in both variants, yet the economically sensible stock position differs by a factor of four or more. What matters is the distribution of basket sizes, not the average parceled into daily buckets.
As the two examples above demonstrate, even for a single SKU, time series flatten critical aspects of the flow. Time series are a lossy representation of transactional history: the projection discards information, and unlike meteorology, what is lost often proves essential. Consequently, the resulting plans are invariably dysfunctional because they ignore critical aspects of the situation. As the weakness lies in the time-series paradigm itself, it is delusional to think the matter can be cured by greater sophistication. Adding more series, extending their horizon backward or forward, increasing granularity, or improving point accuracy won’t fix the problem.
Many more similar examples could be adduced. Even for a single SKU, time series fail: with perishables (each unit has its own shelf life); with repairables (each unit has its own life cycle); with reverse logistics (returns create confusion); with product launches (no series yet), etc. Indeed, it is hard to find a vertical where a sizable share of the business is incompatible with the time-series paradigm—and thus with “planning”.
These single-SKU caveats do not vanish at scale; they explode. In a sizable firm with thousands of SKUs, the time-series lens—and the planning rituals built on it—produces the same errors, only magnified. Consider a sportswear company with its own stores that, among many items, sells a black backpack. Under the planning view, the firm prepares a 12-month sales projection for that backpack and sets production accordingly, then repeats the exercise across the assortment. The practice is so common that few question it, yet its premise is unsound.
The projection is nonsensical. Prominent in-store placement would lift demand; confining the backpack to e-commerce would suppress it. If the company introduces many backpack variants—colors and sizes—the variants will cannibalize demand for the original backpack. If the company adopts an aggressive price, making the backpack the tip of a marketing campaign to draw new customers into the store network, demand will be far higher than if it is merely one unremarkable product within a vast assortment.
The plan doesn’t make sense because each projection is conditional on decisions that have yet to be made, yet the exercise proceeds as if those decisions had already been made. Once the plan is issued, clerks must reverse-engineer their choices to make subsequent commitments appear to conform to it. In practice, the plan is not impossible; it is made “feasible” only through costly approximations—expedite fees, excess buffers, artificial smoothing, and deferred obligations—that keep the ledger aligned while eroding economics. Feasibility here is an accounting artifact rather than an operational truth.
There was nothing special or surprising about Gosplan’s producing infeasible plans. This is a defect inherent in teleological planning. Corporate executives, feudal lords, and political commissars—adopting the same view—would likewise end up with infeasible plans. To understand why, remember that supply chain, at its core, is an economic process: allocating resources that have alternative uses.
When conjuring a plan, managers implicitly command thousands of scarce resources and expect them to be ready at the right time, place, quantity, and cost. The odds of perfect alignment between the plan’s innumerable expectations and the resources actually available to the company—at every time and location throughout the planning horizon—are essentially nil.
Time-series planning—and, more broadly, the teleological vision—does not merely fall short; it misdirects effort. Yet well into the 21^st^ century, it remains the default in most companies and in nearly every supply chain forum. How did a repeatedly refuted notion become orthodoxy? The explanation lies less in logic than in incentives. The next section examines those incentives—the bureaucratic appeal of planning—and shows why the doctrine endures despite its poor results.
Bureaucratic appeal
Corporations are in love with the idea of the strategic plan. They need to pay to figure out where they are going. Yet there is no evidence that strategic planning works—we even seem to have evidence against it. A management scholar, William Starbuck, has published a few papers debunking the effectiveness of planning—it makes the corporation option-blind, as it gets locked into a non-opportunistic course of action. Antifragile: Things that Gain from Disorder (2012), Nassim Nicholas Taleb.
The Gosplan might be gone, but its spirit lives on in large corporations and government agencies. The teleological vision is undeniably appealing, but not solely because of ignorance or stupidity. Its staying power owes less to ignorance than to incentives that reward plan theater; unless those incentives are named and confronted, the doctrine will endure.
At its core, planning is a bureaucratic function executed by back-office specialists. The function is unavoidable: the firm needs a mechanism to allocate a tangled mix of resources across time and locations. They seldom touch the physical flow directly; they orchestrate it indirectly. The allocation problem mirrors the flow it governs and is too intricate for any one person to manage end-to-end. Because the work is analytical and data-heavy, specialists are markedly more productive than generalists at it. Thus, spreading the workload across the entire management layer is rarely viable; the sound pattern is to centralize the task with a small cadre of planners.
As in any bureaucracy, failure is punished—often harshly—while success barely registers. For example, a successful salesman closing twice his expected yearly sales quota is in a strong position to negotiate a steep raise. Yet a planner who halves the inventory write-offs of his peers rarely receives comparable recognition. While a salesperson’s success is viewed as personal, a planner’s achievements are seen as dependent on many other actors inside and outside the firm, such as suppliers. Conversely, a salesperson is seldom dismissed for missing a lucrative deal, whereas a planner who triggers a major write-off may face severe career consequences if the loss is traced to the plan he proposed.
Whether justified or not, such preconceptions pervade large companies, particularly among middle managers. Consequently, members of the planning bureaucracy soon learn that dodging accountability is safer than claiming credit for the occasional success. Predictably, a process-oriented culture displaces a results-oriented one. This outcome is sustained by redefining “good planning” as adherence to codified processes. This twist makes the bureaucracy nearly unassailable: success is redefined as continual compliance with rules that it wrote for itself.
Unfortunately, mere process compliance yields little tangible return in planning. Corporate planning is unlike aviation safety, where each now-rare crash is examined to identify the adjustments that will prevent a repeat. Instead, planning mistakes are so frequent and so significant that such adjustments often introduce at least as many flaws as they fix. As a result, mediocrity remains a long-term characteristic of planning.
In particular, time-series forecasting creates an accountability trap for the planning team. A point forecast is guaranteed to be wrong at the point of use; when shortfalls or overages occur, the miss is visible and, inside the firm, it accrues to the forecasters. Bureaucracies avoid such single-owner exposure. The routine response is to broaden “ownership” of the forecast—pulling sales, marketing, finance, and operations into the ritual—not primarily to sharpen the numbers, but to ensure that any error becomes a collective one.
This maneuver typically appears as a large “collaborative” initiative that claims to improve forecast accuracy. S&OP (Sales and Operations Planning) is one of the most popular flavors of such initiatives. S&OP’s endless meetings ensure that many other parties co-sign the forecasts. There is no empirical evidence that such meetings improve planning quality. In fact, evidence shows that large committees almost always deliver inferior results, whatever the task at hand.
In short, teleological forecasting appeals to clerks because it lets them sidestep both decisions and consequences. Forecasts serve as an occupational buffer that deflects responsibility for the resulting resource allocations. Planners then tweak the forecasting process—often by branding it “collaborative”—to dilute their exposure even further.
Although a straightforward analysis lays bare the flaws of the teleological view, applying that critique to a specific company is often hard. These flaws are often masked by methodological bolt-ons that give the planning process a “scientific” varnish.
Critique is not a method. Firms still need a compass—one that does not pretend tomorrow is a fixed extension of yesterday. The alternative is to treat uncertainty as an input rather than a defect, to regard optionality as an asset to be cultivated and spent, and to judge choices by their risk-adjusted contribution to profit rather than by plan fidelity. This stance, common among entrepreneurs and rare in bureaucracies, recenters agency on allocation rather than on the forecast used to excuse it. We call this the rugged vision, examined next.
The Rugged Vision
The future is irreducibly messy. Customers are fickle, employees capricious, and suppliers unreliable. Yet rivals face the same turmoil. You need not be perfect—only better. Every bump is also a passing lane. Decisive action beats elaborate planning, and luck rewards the prepared. Preparation, then, is the art of cultivating opportunities rather than guessing tomorrow’s exact market.
These remarks distill the rugged vision. The teleological camp treats the future as a journey that demands an ever-more-detailed map—the forecast. If the map is precise enough, no other instrument is needed: it alone should guide the traveler to a known destination. Traveling without such a map feels reckless to its believers.
By contrast, the rugged vision relies on continuous opportunity assessment—a homing beacon, not a map. If that beacon is sound, progress is a matter of moving steadily in its direction. Knowing the destination in advance adds no speed; shifting terrain soon renders static maps useless.
Even though entrepreneurs practice it daily, many intellectuals dismiss the rugged vision. Their scorn springs less from doubts about its efficacy than from moral overtones, as though embracing it reveals a flaw in character.
The first stigma arises from the implicit association of the rugged vision with gambling and other forms of idle speculation. However, while the entrepreneur does take his chances, this isn’t “idle” speculation, and the process is expected to be profitable for both the entrepreneur and his clients. For example, the entrepreneur who launches a product that becomes a massive hit may be judged an undeserved success, as his economic contribution—to the inattentive observer—seems minor. However, to achieve that one success, he may have launched dozens of products, losing money on each that didn’t hit. Furthermore, whether a product is a minor variation or a radical improvement lies solely with consumers. A major commercial success suggests that consumers found something distinctly appealing. Unfortunately, the costs of exploring many angles to better serve the market are largely invisible to outsiders, so success is often dismissed as pure luck—like a lottery win.
The second stigma arises from characteristic behaviors that many observers deem exploitative. Again, observers focus on the benefits—real or imagined—gained by the entrepreneur, skewing their perception. First, in a free market a transaction occurs only when both parties consent. Accordingly, no matter the price asked by the seller, the buyer cannot be “exploited” unless he agrees to the price106. For example, a firm that rushes portable generators into a storm-damaged region may charge far more than the “normal” price. Yet nothing about this is exploitative. Buyers are not forced to buy the generators, and those who do judge themselves better off with one than without107. More generally, in the rugged vision, prices shape the flow: buying low and selling high is its driving force. Yet it is often miscast as exploitative because inattentive observers don’t realize that “buying low, selling high” isn’t luck but requires constant effort and entails unavoidable financial risks as well.
Since the 19^th^ century, intellectual circles have labored to stigmatize the rugged vision108 and have spread this perception to the public. Many quirks of what passes for the “mainstream” supply chain theory nowadays—such as dodging the obvious impact of price on demand—owe much to this stigmatization.
Yet truth is not decided by popularity; the rugged vision is the superior alternative to the teleological one, both conceptually and empirically. Conceptually, the rugged vision embraces the profit-seeking perspective and its attendant traits. For example, adversarial behaviors have no place in the teleological view yet fit naturally in the rugged view. Empirically, the rugged vision proves more resilient to ambient chaos because it is far less dependent on plan correctness, yielding superior risk-adjusted decisions. For example, the very notion of “opportunities” is self-evident in the rugged vision yet missing from the teleological one.
Irreducible uncertainty
The rugged vision assumes that the future is resolutely uncertain, without implying that it is entirely unknowable. The rugged view neither posits that possible futures are unthinkable nor that all are equally likely. On the contrary, this view holds that some futures are much more likely than others, that such futures can be identified, and thus that resources can be allocated to support those probable futures. In turn, customers can be served more diligently and more profitably.
The rugged vision holds that part of tomorrow can be anticipated, often only in coarse, structural terms. Much of what we know is comparative—“a gram of gold will almost surely remain pricier than a gram of silver”—or relational—“SKU A will outsell SKU B at equal price”. Such knowledge is operationally useful yet awkward for equispaced point-time series, which offer only one number per date. Time-series slots cannot express inequalities, orderings, or conditional statements; they flatten relative facts into absolute guesses.
Even when the knowledge is not comparative, its structure resists the time-series mold. Consider aviation: global passenger counts rise in most years, punctuated by rare, severe collapses. A point series forces us to pick one number per future date; it cannot encode “almost always up, except for catastrophic down years with small probability”. Capturing asymmetry and tail risk requires probability distributions or regime variables, not a single baseline path. In short, the time-series formalism is too narrow for the kinds of claims that give firms a real edge.
Seen from this angle, the rugged view is no less “predictive” than the teleological one; it simply uses a different grammar. Rather than filling time buckets with a single baseline path, it assigns probabilities to families of outcomes and records comparative facts—e.g. that SKU A will likely outsell SKU B at equal prices, that a heat wave lifts water sales, or that lead times carry a fat right tail. What matters is not a solitary number per date but the priced set of alternatives these statements unlock. A forecast is deemed “good” insofar as it creates an economic edge—enabling the firm to commit, defer, hedge, or reprice better than rivals—rather than because it minimizes an academic error score.
Unlike the teleological view, which flattens tomorrow into a single quantity-over-time, the rugged view treats the future as a portfolio of uncertainties spread across the whole flow. Quantity demanded is only one variable. Prices and commercial terms can shift; a rival may extend the warranty from three to five years, forcing a response; product mix and substitution evolve with merchandising. Lead times wobble—often with fat-tailed delays; purchased and operating costs drift; suppliers and production lines sometimes fail. Each of these deserves its own probability law because each reshapes opportunity costs as much as demand does. The practical task is to price these uncertainties jointly rather than pretend one scalar forecast can stand in for them all.
In the rugged view, there is not one future but a boundless multitude. Reducing that multitude to a handful of scenarios—let alone a single forecast—is wishful thinking. Because the future is only partly knowable and brimming with uncertainties, listing every plausible outcome is impossible. Any effort to craft a sweeping “grand-master plan” therefore borders on self-deception.
Instead, the rugged view treats planning as an attempt to reconcile diverse and contradictory insights about the future in order to allocate resources most profitably. Going big on one unlikely outcome may be acceptable if the returns are high and offset the risks. Conversely, a very likely outcome may merit no allocation if no profit can be had, however certain.
In summary, the rugged vision treats tomorrow as a portfolio of heterogeneous clues. It refuses to compress them into a single time-series baseline and keeps each in its native form—probabilities, regimes, inequalities, and comparative statements. The practical problem is synthesis: turning disparate clues into consistent choices. That task belongs to the tooling, not the vision. The remainder of this chapter introduces the theory and instruments that perform this synthesis.
Before the machinery, one principle: in the rugged view, the decision—the allocation itself—comes first.
Opportunistic allocations
The rugged vision treats the future not as a fixed “baseline” of units to purvey, but as a rolling stream of opportunities—some dismally unprofitable, others very lucrative. Every decision is, above all, an economic valuation of the opportunity to be seized.
In the rugged vision, the decision sits atop—a deliberate allocation of scarce resources to an opportunity. Forecasts are inputs to that choice; in the teleological vision, by contrast, the forecast wears the crown. This sharp, distinctive focus complicates any attempt at dialogue between proponents of the two visions. To the rugged exponent, forecasts are sophisticated but ultimately secondary, wholly trumped by the need to act decisively. To the teleological exponent, the forecasts and their associated plan are inseparable from a rational course of action. Decisions not grounded in a plan are seen at best as shortsighted and at worst as whimsical.
Yet while the rugged vision is fundamentally opportunistic, it is just as forward-looking as the teleological vision. Indeed, every allocation of resources comes with an unavoidable opportunity cost. As a result, a seemingly profitable opportunity might be rejected because it could open the door to a better one or, conversely, spawn problems that erode the expected gains.
For example, let’s consider an airline leasing an expensive, repairable aircraft part to a peer so that one of its jets is not grounded for lack of that component109. As the part is urgently needed, the airline can command a very high price for the lease from its peers. Yet the airline will frequently decline to lease the part, even if, by basic accounting rules, the operation appears excessively profitable on paper. Indeed, the airline must consider its own fleet. When leasing the part, the airline increases the risk of grounding one of its own aircraft precisely because this very part ends up missing. The cost of this risk can entirely offset the profit from the lease, despite a price point that, on paper, appears highly profitable.
Let’s note that the teleological vision is entirely sidestepping the matter of opportunity costs. Indeed, from this perspective, either the plan is deemed feasible—in which case no customer will be left unserved—or it is deemed infeasible, in which case demand forecasts are revised and lowered until the plan becomes feasible again. In the teleological vision, the very notion of opportunity is buried—along with nearly every other supply-chain concern—inside the time-series projection of future demand.
Exact calculation of these opportunity costs is difficult in practice and acknowledged as such in the rugged vision. Thus, as an alternative to explicit mental arithmetic on opportunity costs, the rugged vision favors comparative assessments. For example, should the company allocate its last unit to a new but unknown customer, or keep it implicitly reserved for a loyal customer the company isn’t willing to disappoint? While comparative assessments can be made through sophisticated methods, the rugged vision emphasizes a general preference for looking at every allocation through the lens of its alternatives.
Through such comparative assessments, one can also decide whether to allocate at all: if a preferable deferred opportunity exists for the same allocation, the current one should be disregarded. For example, the company can stop buying additional units, reserving cash for later, once current stock is further depleted and needs are clearer.
As a result, the rugged vision doesn’t need a “plan” in the sense of the teleological vision. Resources will be allocated incrementally until opportunity costs are deemed too high to continue. The existence of preferable deferred opportunities is all the company needs to avoid exhausting its resources through premature allocations. Thus, despite its opportunistic stance, the rugged vision is no more short-sighted than the teleological one. Both rely on estimates of the future and are therefore forward-looking, even if their predictions differ in form.
Once allocations are comparable, they can be prioritized. A single, continuously resorted priority list can steer the entire flow as new information revises each valuation. At enterprise scale, this list easily runs to millions of entries—each admissible allocation—well beyond human curation. The rugged vision says what to rank and why. How to maintain and act on that ranking is an engineering matter we will address when introducing the instruments that mechanize the vision at scale. First, we need one more ingredient: the recognition that the future is not a mirror of the past but the consequence of decisions yet to be made—the asymmetry of time.
Asymmetry of time
Unlike the perspective common in the natural sciences—where many laws are effectively time-reversal symmetric and tomorrow is treated as another stretch of the same process—the rugged vision treats the future as shaped by choices not yet taken. In business, tomorrow is not a mirror of yesterday but the consequence of present commitments. Forecasting while deferring the decision reverses causality: decisions cause outcomes, not the other way around.
We have already discussed the case of a sportswear company selling backpacks. Future sales cannot be thought of as a mere projection of past sales, because the company can shape demand along many dimensions: lowering the price, introducing a larger backpack model, or offering it in a new color, paying for advertising, featuring the item prominently in stores, etc. Thus, from the rugged vision, it does not make sense to treat the coming year as an extension of the previous year. The coming year is categorically different from the previous year.
Even more unsettling to teleological planners, the rugged vision treats dependence on decisions not yet taken as perfectly normal. Present commitments may be made expressly contingent on later choices: we secure and price options today precisely so that we can decide with fresher information tomorrow, when opportunity costs are clearer. To the rugged practitioner, this is not circularity but prudence—acting now to enlarge the future choice set, while deferring selection among those choices until the market has revealed more.
For example, a fashion brand might negotiate and pay for a large upfront capacity commitment from a trusted supplier. This commitment puts the supplier in a position to aggressively source the raw materials, such as textiles and leather. This mitigates the risk of the supplier running into delays caused by later difficulties sourcing the raw materials, while also paving the way for lower acquisition costs. However, when the commitment is struck, the brand has not yet decided what it will order. Under the rugged vision, those details can wait. Provided the capacity is broadly right, postponing the purchase order is preferable because it preserves the option to align later orders with the market’s latest desires.
The value of preserving options is another cornerstone of the rugged vision. Indeed, whenever the company commits to a course of action, it loses some of its agency. For example, once a manufacturer triggers a production batch, the raw materials are consumed and can no longer be used for an alternative production run. Similarly, if the same manufacturer commits to triggering the production batch next month, the matching raw materials are effectively reserved for that upcoming production.
Beyond direct commitments, options also erode through precedent—the ratchet effect. Once a behavior is observed, counterparties update their expectations and make reversal costly. A fashion brand that ends a season with deep markdowns trains customers to delay purchases and expect the same discounts next season; a “temporary” service upgrade becomes the new baseline; a one-off MOQ waiver becomes the new rule. Like a mechanical ratchet, motion in the tightening direction is easy, while release is resisted. In the rugged view, every action is priced not only for its immediate margin but also for how it expands or shrinks tomorrow’s option set; moves that lock future choices must clear a higher bar than those that preserve flexibility.
More generally, whenever part of the future is locked in, it becomes functionally identical to the past, since it can no longer be acted upon, even if events have not technically unfolded yet. Not only does the rugged vision treat the future as radically distinct from the past, but it also cherishes preserving what makes the future distinct—namely, the company’s agency. To rugged practitioners, committing oneself to a final course of action is a last resort, undertaken only when the benefits of commitment outweigh the opportunity costs.
For those who hold the teleological vision, this stance is often perceived as fickle or mercurial. There is no commitment to one’s former self, only to formal agreements involving third parties. For example, when probed about a future course of action, an exponent of the rugged view will simply offer his best guess, without any particular attachment to the statement. It’s just one option—tentatively the most likely—but there is no certainty. This option coexists with a vast number of alternatives that, in aggregate, are even more likely. Later—as events unfold—the course will be revised in light of new information.
From the teleological viewpoint, the existence of an underlying plan is taken for granted: it is the central characterization of the future. The possibility that there is no plan at all is nearly unthinkable and is treated with contempt as amateurism when encountered. Thus, changing the course of action implies changing the plan. Such an outcome is seen as a reflection of an underlying problem. This outcome reflects either inadequate execution—failure to stay compliant with the plan—or an inadequate plan in the first place110.
Quantifying a future deemed dependent on decisions yet to be made yields numerous technical difficulties. In particular, time-series forecasts and most planning instruments of the teleological vision end up being inadequate. As the supply-chain literature offers next to nothing to support the rugged vision, most of its exponents rely on back-of-the-envelope calculations. Those calculations are deemed crude by teleological planners. Yet, as will be shown in what follows, instruments supporting the rugged vision can match—indeed, often exceed—the sophistication of classic planning methods.
Superiority of the rugged vision
While the rugged vision enjoys virtually no support from intellectual circles111, it is vastly superior to the teleological vision. Empirically, when successful entrepreneurs disclose their vision, they invariably lean toward the rugged vision. Ingvar Kamprad (Ikea), Jeff Bezos (Amazon), and Warren Buffett (Berkshire Hathaway) are a few who took such public stances. Successful entrepreneurs devote nearly all their effort to running their businesses and thus do not produce as many papers or books as intellectuals who specialize in such work. Hence, writings favoring the rugged vision are rare.
Furthermore, there is ample empirical evidence of the dismal results of the teleological vision. For seven decades the Soviet Union’s Gosplan imposed severe economic inefficiencies on more than 100 million people. More broadly, in both private and public sectors, large multi-year plans frequently derailed, generating immense costs. Conversely, grand-planning successes that made a large company’s fortune are unheard of. At best, plans execute with limited overhead—in cost and delay—and that is the full measure of “success” here.
Those observations—both successes under the rugged vision and failures in its absence—stem largely from a single factor: how uncertainty is managed. The teleological vision tries to eliminate uncertainty; the rugged vision embraces it. Thus teleological methods are fragile to shocks, whereas rugged-vision methods can mitigate—or even profit from—the same shocks.
In a world with little or no uncertainty, the teleological vision might perform admirably. We do not live in such a world. Free markets alone suffice to create substantial uncertainty for all participants. Indeed, free markets enormously magnify minor variations. For example, Walmart drove competitors into bankruptcy—Kmart in 2002 and Sears in 2018. Walmart took Kmart’s revenue from over $36 billion in 1994 to zero despite being only 2.2% cheaper112. In a free market, being 1% worse than a competitor typically yields a 100% loss of market share within a decade or two. Thus, even if non-market sources of uncertainty—wars, hailstorms, earthquakes, epidemics—were vanquished, irreducible uncertainty would still define free markets.
Supply-chain instruments—whether algorithms or methodologies—that support the rugged vision are scarce—an unsurprising consequence of its lack of supporters among intellectual circles. Most practitioners rely on experience and the intuition it yields. Yet nothing prevents us from introducing such tools—an endeavor these chapters undertake step by step.
Predictability of human affairs
A supply chain’s performance hinges on the capacity of the company, both its managers and its software systems, to adequately anticipate future events. These events include, above all, human decisions. Perfectly predicting natural events is conceivable—though rarely attainable beyond simple cycles like tides—but perfect foresight in human affairs is impossible. All it takes is a contrarian customer113 changing his mind to undermine a supposedly “perfect” forecast. Yet habit makes actions somewhat predictable in the short run.
For instance, when a supermarket manager places a pack of yogurt on the shelf with a price tag, he expects a customer to buy it. Moreover, he expects the sale to occur well before the yogurt reaches its expiration date. Accurate anticipation keeps write-offs and stockouts in check—a discipline most supermarkets master. Profitable supermarkets prove that such anticipations can be produced at scale as routine business.
In general, supply chains remain profitable because managers anticipate future events well. For the most part, this means anticipating decisions to be made by others. Until the late 19^th^ century, the process relied on intuition and experience. In the early 20^th^ century, formal analytical methods appeared, and model accuracy began to be measured rigorously. These methods flourished after 1950, buoyed by steadily increasing computing power. This ushered in the time-series paradigm, which, as discussed previously, became the dominant predictive approach in supply chain.
Yet the time-series lens is brittle in business settings. It abstracts away the very levers that make tomorrow unlike yesterday and compresses sparse, discrete events into fractional averages. Where it helps, it does so for two reasons only: short-run momentum and calendar cyclicities. We examine those narrow strengths, then replace the lens with instruments that fit the economics of the flow: (i) probabilistic forecasts that attach distributions to demand, lead times, prices, and returns; (ii) high-dimensional forecasts that capture baskets, substitution, and cross-effects across items, channels, and locations; and (iii) functional (policy-conditioned) forecasts that speak in the native objects of the flow—orders, allocations, and prices—rather than in evenly spaced time buckets. These tools do not guess “the” future; they price alternatives and make opportunity costs explicit.
Momentum and cyclicities
Time-series forecasting treats human affairs as periodic measurements projected forward. Once past numbers are recorded, the task is to propose sensible values for future dates—and to do it as faithfully as our tools allow.
With millions of papers114 on time-series forecasting, it is difficult to overstate the perceived importance of this paradigm in academia. Yet, for a corpus of this magnitude, it has surprisingly little to offer. It rests on two principles that give time-series models predictive power: momentum and cyclicities. They matter, but their limits are obvious.
Momentum115, in supply chain settings, is the short-run reluctance of people to change their habits. This reluctance is not mere stubbornness; change often carries sizable costs. Assessing whether change improves on the status quo may require considerable effort, and the rollout may undermine the value of existing productive assets.
For example, a manufacturer may decide to source its machine tools from a lower-priced supplier. As machine tools are not simple commodities, assessing suitability is time-consuming. Teams must be retrained and processes adjusted to make the most of the new machines. Finally, existing spare parts may lose most of their value if they are incompatible with the new supplier’s machines.
Thus momentum explains why the moving-average model family—including exponential smoothing—often provides a decent short-term baseline: the near future resembles the recent past. Moreover, momentum governs more than demand: lead times, return rates, defect rates, purchase and selling prices, etc.
As habits shift, momentum’s predictive power wanes; the Lindy effect tells us how fast.
If a book has been in print for forty years, I can expect it to be in print for another forty years. But, and that is the main difference, if it survives another decade, then it will be expected to be in print another fifty years. This, simply, as a rule, tells you why things that have been around for a long time are not “aging” like persons, but “aging” in reverse. Every year that passes without extinction doubles the additional life expectancy. This is an indicator of some robustness. The robustness of an item is proportional to its life! Antifragile: Things That Gain from Disorder (2012), Nassim Nicholas Taleb.
The Lindy effect governs habits as surely as objects, and in a shifting world it is critical in managing allocation risk. Substantial investments demand momentum that is both strong and durable. Practitioners use the Lindy effect to assess the lifespan of a given momentum.
For example, a two-week surge in breakfast-cereal sales might suffice for a supermarket to expand its inventory. However, the brand observing the same surge might delay capacity expansion by several months. The supermarket’s bet is modest; if the surge is transient, the overstock will work off over time. By contrast, the brand’s capacity investment is far larger and, if the surge fades, must be written off.
Cyclicities capture the human tendency to anchor decisions to the calendar. In practice, the usable calendar regularities are few and stable: hour of day, day of week, day of month, and day of year. Everything else of lasting value is a variant or composition of these four, plus a handful of quasi-cyclical events that do not track the Gregorian calendar exactly.
Hour of day. Stores have opening hours; e‑commerce has order cut‑offs and late‑evening spikes. Sub‑daily ledgers reveal sharp intraday patterns (e.g., lunch‑break lulls, end‑of‑day surges). With daily data, intraday cycles mostly vanish from view except through induced day-of-week effects—for example, Sunday closures zero out retail demand.
Day of week. Households shop on weekends; B2B demand concentrates Monday through Friday; returns cluster on Mondays. For many assortments, the weekly profile explains more variance than any other single factor.
Day of month. Payroll, benefits, bill runs, and end‑of‑period targets create month‑end “hockey sticks”—especially in shipments. Note the distinction: a month-end shipping spike is usually a bureaucratic gating artifact, not genuine consumption. Treat it as a calendar effect on observed flows rather than a persistent lift in underlying demand.
Day of year. Seasons, vacations, and civic holidays dominate the annual profile. “Seasonality” here is no mystery; it is the calendar’s imprint on habits (heaters in winter, sandals in summer). Retail “events” are likewise calendar-bound; Black Friday (the day after the fourth Thursday of November in the USA)116 reliably realigns baskets every year.
Beyond the four base frequencies sit a few quasi‑cycles: religious or lunar events (Ramadan, Chinese New Year, Easter) that drift relative to the Gregorian year and must be handled with explicit, locale-specific calendars rather than naive monthly indices.117 School calendars and local festivals are of the same kind: discrete calendars with predictable date lists, not free-floating “patterns” to be discovered.
From a modeling standpoint, the most economical treatment is a small set of multiplicative indices—one per relevant calendar factor—regularized toward $$1$$ under weak evidence. Geography and channel merit separate indices (what moves on Saturday in retail often does not in B2B); closures imply structural zeros rather than imputation; sparse SKUs borrow strength from their peers through hierarchical pooling. This keeps the calendar explicit—and portable—across assortments.
Put plainly, once momentum (the short‑run persistence of habits) and a compact calendar (the four base cyclicities plus a few explicit quasi‑cycles) are in place, almost all of the practical signal that time‑series models can harvest has been harvested. Elaborate decompositions—Fourier fireworks, exotic lags, baroque state machines—rarely survive out‑of‑sample in business settings. Everything beyond momentum and calendar belongs to other instruments (pricing, promotion, substitution, lead-time modeling), not to the time-series lens itself.
The next section shows that, after momentum and calendar are accounted for, the remaining tail-chasing within the time-series paradigm yields vanishing returns for supply chains.
The dead end
Despite tens of thousands of new papers each year, time-series forecasting has proved a technological dead end for supply chains118 for years. The clearest demonstration of this claim appears in the M5 results, a large public forecasting competition held in 2020—the largest to date—using a store-level sales dataset from Walmart.
In this competition, a team from Lokad—a supply chain software specialist—placed first at the SKU level119—the most disaggregated and, from a supply chain perspective, the only level that ultimately matters for flow decisions. Their winning forecasting model120 was a simple multiplicative parametric model, using only momentum and cyclicities as introduced in the previous section. Furthermore, the vast majority of those parameters were basic cyclicity indices, i.e., multiplicative coefficients reflecting the “day of the week”, “day of the month”, and the “month of the year” factors.
At the SKU level, this model outperformed more sophisticated approaches, including deep learning models (with millions of parameters) and gradient-boosted machines (with thousands of trees). At other aggregation levels, this simple model remained within a narrow margin of the winners’ accuracy.
With 909 teams worldwide in the M5, every notable time-series method was surely tried. Yet, in the end, a simple model achieved comparable—if not better—accuracy than its more sophisticated counterparts. The reason is simple: no model managed to capture valuable information from the historical time series beyond momentum and cyclicities. No model managed this feat because such information is not present in the M5 time series.
For the Lokad team, far from being a surprise, this outcome confirmed a realization formed about a decade earlier, after running numerous time-series forecasting benchmarks on datasets from dozens of corporate clients. These datasets hide no undiscovered cyclicities, no latent lags where one series predicts another, and no secret trends waiting to be found121. More generally, within business records, there are no alternative patterns to exploit through more sophisticated time-series forecasting models.
The absence of evidence isn’t evidence of absence. While alternative time-series patterns—i.e., neither momentum nor cyclicities—remain elusive, such patterns may yet be discovered. Yet the odds of a breakthrough after such a staggering volume of research are low. By the end of WWII, the predictive power of momentum and cyclicities had already been identified. The M5 competition then illustrated that, 70 years and millions of papers later, no further principles have been uncovered. Most of what has passed for “progress” since then has merely refined the parameterization of those two effects—smoother trend terms for momentum and ever-finer calendar indices for cyclicities—yielding marginal gains with steep diminishing returns once a clean multiplicative baseline is in place.
There are numerous patterns to exploit beyond momentum and cyclicities, but they lie outside the narrow confines of the time-series paradigm. Thus, to take advantage of those patterns, the time-series paradigm must be abandoned.
The persistence of ever-new time-series “advances” owes less to misunderstanding than to incentives: novelty markets better than admitting that, beyond momentum and cyclicities, little signal remains to harvest. Hence the parade of variations decade after decade, each adopting whatever technique happens to be fashionable in computer science122. For supply-chain forecasting, these waves add little of substance, as the M5 results above illustrate.
The law of small numbers
Momentum and cyclicities, as predictive principles, are undeniably powerful. When large numbers are involved, human affairs become quite predictable using nothing but those two principles. For example, traffic-jam dates in France can be predicted years ahead with near-certainty. Those traffic jams reflect tens of millions of individual decisions entrenched in lifetime habits. Transport infrastructure is still evolving, but the progress is slow enough not to interfere much with such predictions a few years out.
Unfortunately, supply chains are fundamentally about small numbers. Whenever a number characterizing the flow gets large, say exceeds 20, batching occurs and the number is lowered, owing to economies of scale along the way.
Let us illustrate this principle with a local shipment of small mechanical parts. A single urgent part goes by scooter. As quantities rise, the mode changes in discrete steps: a few dozen parts are shrink-wrapped into one bundle and ride on the next courier run; several bundles are packed into a carton and sent by delivery van; a stack of cartons becomes a pallet moved by a medium-duty truck; a handful of pallets justify scheduling a full-truckload. In practice, the flow advances in quanta—unit → bundle → carton → pallet → truckload—rather than one unit at a time.
As a result, supply-chain time series are almost invariably sparse (observations are mostly zeros) and erratic (non-zero observations vary significantly).
Time-series methods perform poorly on sparse series: they yield fractional daily volumes (e.g., 0.3 units/day) that lack operational meaning at the point of use and, unsurprisingly, prove inaccurate against real decisions. A common palliative is to “hide” sparsity through aggregation—lengthen the time bucket (days $$\rightarrow$$ months) or broaden the scope (SKU $$\rightarrow$$ category).
Yet even if aggregation technically yields more accurate time-series forecasts, it is little more than an optical illusion. The decision’s structure—resource allocation—dictates the prediction’s granularity, not the other way around.
For example, consider a store manager selling a few jars of a given mustard brand each month. In his case, the replenishment decision has two fields: a date (when to reorder) and a quantity (how many units). The manager needs to know if today is a good day to place the reorder, not whether this month is the right month to reorder. Similarly, he needs to know how many jars (same brand, same size) to reorder. An assessment in condiment units—the product category—is useless to the manager.
The law of small numbers explains why the hope of engineering precise time-series forecasts is greatly misplaced—even considering situations that aren’t entangled in decisions that have yet to be made. Small numbers render the data irreducibly erratic. This is not a transient defect of present-day statistical instruments, soon to be fixed by superior ones. On the contrary, whatever numerical instruments we bring to supply chain must be tailored to small numbers and otherwise limited datasets.
These constraints do not call for a finer time grid or a more elaborate smoother; they require a different grammar. When counts are small and actions are discrete, the meaningful unit of prediction is not “one number for next Tuesday” but a probability law over admissible outcomes—how many, if any, and how late. The first instrument in that grammar is the probabilistic forecast: a distribution that accommodates sparsity, integer quantities, and fat-tailed surprises, and applies equally to demand, lead times, returns, and prices. We start with lead times and demand to fix ideas.
Probabilistic forecasts
Probabilistic forecasting considers all possible future outcomes for an event and assigns probabilities to each. A probabilistic forecast is the complete set of those probabilities. A non-probabilistic (classic) forecast that pinpoints one outcome is a point forecast; the term is rarely used, since “forecast” usually implies “point forecast”. Probabilistic forecasting is fundamental for supply chain because it embraces the law of small numbers, both theoretically and empirically. It marks our first major departure from the equispaced point time-series paradigm.
Probabilistic forecasts quantify uncertainty across the entire flow, not just demand. This is our second break with mainstream practice: we attach distributions to every uncertain driver that moves cash and atoms—lead times, inbound fill rates, supplier reliability, purchase and selling prices, returns, scrap, and available capacities. Throughout this book, “probabilistic” is used in this broad sense. To fix ideas, consider varying lead times. Neither supplier nor carrier is perfectly reliable; unscheduled delays happen. What we need is the distribution of the time-to-receive for a just-placed order. A probabilistic lead-time forecast provides exactly that distribution. The histogram below summarizes the day-by-day probabilities.
The histogram shows that delivery in under three days has zero probability. This reflects the selected shipping method’s minimum transit time. The distribution’s mode occurs on the third day, where the probability is highest. This is the nominal lead time. It is the duration expected when both supplier and carrier perform as planned. Small but non-zero probabilities stretch far to the right; they represent disruptions that inflate lead times well beyond the nominal. In the worst case, the delivery never occurs.123
Like point forecasts, probabilistic forecasts can be computed from historical data. Setting modeling technicalities aside, let us clarify what probabilistic forecasts bring to supply-chain decision-making.
Probabilistic forecasts convey richer information than point forecasts. However, they are not intrinsically more accurate. Probabilities add an information dimension absent from point forecasts. In particular, one can collapse a probabilistic forecast into a point forecast by taking the distribution’s mean or median. Thus, a more accurate probabilistic forecast will mechanically yield a more accurate point forecast. Given a probabilistic model and a point model, empirical validation is required to determine which is more accurate. As classes, “probabilistic” and “point” models confer no systematic accuracy advantage. Nevertheless, the extra information provided by probabilistic models is often critical for supply chain decisions.
Revisiting the varying-lead-time example above, note that the distribution’s fine print is critical for managing late-delivery risk. Assuming goods arrive at the median lead time would be a grave error: by definition, half the shipments would arrive late. Conversely, basing operations on a far-right value of the distribution—a high quantile—is equally unreasonable. Most goods would arrive much sooner, creating problems such as excessive inventories.
Lead times are only one of many examples where variability and uncertainty critically impact supply chain decisions.
Revisiting the earlier single-buyer variant from Limits of planning. Suppose a factory has shipped 1,000 units per day for years to a single customer. In time-series terms, the point forecast is a flat line at 1,000. Now add information the time series ignores: the account is long-lived but not immortal—the weekly churn probability is about 0.1%. One year out (52 weeks), the expected daily demand is $$(1-0.001)^{52}\times 1000 \approx 934$$ units. Mathematically correct, perhaps, but operationally nonsensical. A year from now, daily demand will be either 1,000 or 0.124 Any capacity plan based on “934 units per day” will misallocate resources.
A probabilistic forecast, by contrast, makes the collapse risk explicit: most probability mass remains on “still 1,000”, with a smaller mass on “drops to 0” over the horizon. That view drives a two-pronged policy: secure current service while the account persists, and prepare a contingency for its eventual loss. The point forecast attempts to predict a single “true” future and averages away risk; the probabilistic view prices alternatives and prevents the planning failure the flat line invites.
Thus, probabilistic forecasting reveals things that point lenses cannot show. Attempting to capture a “one true future”, the point forecast remains blind to otherwise obvious patterns. Moreover, the probabilistic perspective typically permits leveraging additional data beyond the narrow historical time series, or extracting more signal from the same data. We illustrate this point with another example.
Consider a DIY store with a product that sells on average 0.5 units per day. Although unrealistic, for clarity assume demand is constant with no cyclicity. The store is replenished daily from a warehouse that serves similar stores. The demand forecast is again a flat series of 0.5 units per day. Now consider the probabilistic forecast for daily demand, illustrated below.
The product is indeed a corner protector intended for childproofing. Although protectors can be purchased individually, most customers buy them in multiples of 4. The histogram above shows this: the probabilities of daily demand at 0, 4, 8, and 12 units are much higher than for other counts. This probabilistic demand forecast makes clear that the store should hold 8 units in stock — perhaps 12 — to deliver high service quality. This stock level could not have been anticipated by merely considering an average demand of 0.5 units per day.
Probabilistic forecasts not only embrace the law of small numbers but also eliminate the need to handle nonsensical fractional projections. While 0.5 units per day may serve as a convenient abstraction to summarize demand for a packaged product, customers cannot buy fractional units: only integer quantities are admissible. Many authors and software solutions shrug off this issue, but that attitude is misguided. Stock can only be replenished in integer quantities. Accepting a fractional point forecast introduces large approximations that can occasionally generate substantial overheads for the company.
For example, if cycle demand is intermittent—say 0.1 unit per period—there is a real risk it will be rounded to zero. When this happens, no inventory is replenished; the observed demand disappears and the flow halts. Far from a theoretical edge case, this situation—an inventory freeze—is a routine problem for inventory policies governed by point forecasts rather than probabilistic ones. More generally, probabilistic forecasts naturally match the dominantly discrete nature of the flow, a consequence of the standardization principle discussed earlier.
Although any distribution can be reduced to a single number—mean, median, or mode125—doing so discards the information that gives the probabilistic view its value. The benefit lies in the entire shape—its asymmetries, tails, and multimodality—not in a hand-picked “central” value. No single summary can preserve that information. The next chapter shows how to make decisions that exploit the full distribution rather than collapsing it back into a single point.
Generating probabilistic forecasts is arguably more technical than producing point forecasts; however, in this age of powerful computers the overhead is largely inconsequential, provided appropriate methods are used. Because of this hurdle, one might be tempted to do the reverse: first compute a time-series point forecast, then convert it into a probabilistic forecast. For example, one may approximate a sum over a segment with a normal distribution, as is sometimes done for safety stocks.
Unfortunately, this approach yields none of the intended benefits. Probabilistic forecasts are superior because they carry more information, but that extra information does not appear out of thin air. That extra information is extracted by processes specifically designed to draw more information from historical data than methods aimed at producing point forecasts.
A further subtlety: a probabilistic forecast is not vindicated merely because the realized outcome had non-zero probability under the distribution. What matters is the weight the model assigns to the realized outcome. A good model concentrates probability where reality tends to fall and withholds it elsewhere. A bad model does the reverse—sprinkling tiny probabilities everywhere or assigning 1% to events that occur weekly. Thus probabilistic forecasts should be judged not by binary hits but by how much probability they allocate to what actually occurs, while remaining as concentrated as the evidence permits. This calls for proper scoring rules that reward well-calibrated, sharp distributions and penalize misplaced confidence.
Accordingly, we assess probabilistic models using metrics tailored to distributions. Although probabilistic metrics—e.g. log-likelihood—are less familiar than point-forecast counterparts—e.g. mean absolute error—they are not mysterious. The fine print of these metrics is technical and not critical to the present discussion. For now, we can proceed knowing those metrics exist and that, given two probabilistic models, they indicate which model is deemed “best”.
Fat tails
In the natural sciences, many phenomena follow normal (Gaussian) distributions — a remarkable fact. For example, across the world’s population, men’s heights are approximately normally distributed. Similarly, men’s weights are approximately normally distributed. Even if we filter the sample—for example to blue-eyed men or to men living north of the 45th parallel—the distribution remains approximately normal. We obtain similar normal distributions for body temperature, bone density, or sleep-cycle durations.
Normal distributions are ubiquitous across the natural world. Air temperatures, rainfall, and wind speeds over weekly or longer horizons follow normal distributions. They also appear in deviations of apparent stellar brightness and in gene-expression levels across human and nonhuman populations.
As mathematical tools, normal distributions are exceptionally convenient. They benefit from simple analytical expressions. They enjoy simple closed forms, ubiquitous open-source libraries, and a rich algebra: sums of normals remain normal, and appropriately scaled sums of many random variables converge to a normal, etc.
Unsurprisingly, most supply-chain authors embrace normal distributions; yet the view is fundamentally wrong. Once human decisions enter the picture, thin tails are the exception. Most business-relevant variables are fat-tailed: extreme outcomes occur far more often than the bell curve allows; distributions are typically skewed; tails dominate totals. Modeling them as normal systematically understates risk and misprices opportunity.
Consider a stadium with 10,000 spectators. No single person can account for even 0.1 % of the crowd’s total body mass; human weight is thin-tailed as a matter of biology. Net worth is the opposite: one attendee may plausibly own more than the wealth of everyone else combined. Wealth is fat-tailed126.
More generally, anything akin to entrepreneurial success yields fat-tailed outcomes. The top 10 smartphone brands sell more units than all other brands combined. In any given year, the top 100 musical artists worldwide sell more songs than all other artists combined. In French presidential elections, the top three candidates have always received more votes than the two dozen candidates that trail them.
Supply-chain flows are likewise prone to fat-tailed shocks: demand can jump to implausible levels with little warning. In December 1973, a throwaway joke by late-night host Johnny Carson on The Tonight Show about an impending toilet-paper shortage triggered a nationwide run; shelves emptied not because supply changed but because expectations did. Nearly five decades later, the same SKU showed the same tail from a different cause: in March 2020, amid lockdown uncertainty, households stockpiled staples and U.S. toilet-paper sales spiked roughly eightfold127—a dramatic swing for an item that is usually flat. Different triggers—rumor versus policy shock—produced the same operational reality: a long, thin tail became a sudden, overwhelming wave.
Sales can also grow markedly as a result of deliberate marketing campaigns. For example, with its “Just Do It” campaign and strong product, Nike lifted its domestic sports-shoe share from 18% to 43% and grew worldwide sales from $877 million to $9.2 billion over 1988–1998. By 2021, the figure reached $44.5 billion—a roughly fiftyfold increase over a few decades.
Engineering feats can likewise leave one company utterly dominating its market. In 2024, SpaceX placed about 90% of the world’s tonnage into orbit. Even more remarkable, it achieved this with only a tiny fraction of the investment made by its competitors—almost entirely government programs—to build orbital launch capabilities.
Demand is not the only phenomenon subject to fat tails; most flow variables are as well. Prices can swing to astounding levels—in both directions. For example, on April 20, 2020, the price of West Texas Intermediate crude fell below zero for the first time, reaching $-37.63 per barrel. The negative price resulted from expiring futures: holders were obligated to take physical delivery, storage was full, and they dumped contracts at any price to avoid logistical and financial consequences.
Lead times also exhibit extreme behavior. As of 2023, Airbus reported a backlog of over 7,000 aircraft—roughly 7–8 years of output at current capacity. High demand for fuel-efficient models such as the A320neo family has significantly contributed to this backlog. The backlog also reflects the high unit capital cost and lengthy production times. Moreover, other notable airframers were suffering comparably large backlogs.
From a supply chain perspective, fat tails matter because profits and losses are dominated by extreme events. Far from being unpredictable, ubiquitous fat-tailed distributions guarantee that extreme events will routinely appear. Thus, probabilistic forecasts are not sufficient; models must reflect the fat-tailed nature of supply chain phenomena.
Conversely, any supply-chain model that adopts normal—or otherwise thin-tailed—distributions is guaranteed to understate risk and, consequently, to produce fragile decisions. Such fragility puts the firm’s survival in jeopardy: markets—especially when more prudent or opportunistic competitors exist—punish it when extremes arrive.
The presence of normal distributions in supply chain materials—whether publications or software—should be treated as a litmus test for incompetence. This includes all thin-tail methods, such as safety stocks. Indeed, a casual look at empirical distributions from real-world supply chains shows the normal distribution is nowhere close to a reasonable fit128. Thus, by advocating such methods, authors reveal they have not even bothered to look at any real-world supply chain. A company embracing this nonsense, in any guise, only does its competitors a favor.
The price of knowledge
Probabilistic forecasts braid together two distinct kinds of uncertainty; separating them clarifies both modeling and action.
For overseas lead times, weather systems and port congestion inject variability no party can eliminate; even with perfect information and infinite history, the day a container clears the terminal still varies. That irreducible variability is aleatory.
By contrast, when the firm has placed only a handful of orders with a new supplier, the shape of the lead-time distribution is poorly known. Each additional order, clearer timestamps, and tighter Incoterms shrink that ignorance. This is epistemic uncertainty—uncertainty about our model of the phenomenon, not the phenomenon itself.
he distinction matters because, in supply chain, past observations used to generate point or probabilistic forecasts do not fall from the sky. They result from past decisions—that is, from past allocations of resources. Thus, knowledge about the future is paid for. For example, a retail network may place an extra product on the shelves of selected stores to assess demand potential. This is an investment—both the cost of goods and the shelf space—for uncertain returns.
Investing resources to reduce epistemic uncertainty—that is, to acquire knowledge—is fundamental in supply chains. To those who hold a rugged vision, continuously investing scarce resources to probe the market is so self-evident it scarcely merits mention. It is simply the natural way of conducting business.
Yet for mainstream supply chain authors — who mostly adhere to the teleological view — this entanglement between allocations and knowledge is denied rather than merely left unsaid. While most authors realize that market knowledge does not appear out of thin air, they turn to “market studies” and similar bureaucratic instruments129 to acquire knowledge beyond what historical data offers.
From the teleological perspective, the chief benefit of the market study is that it leaves the planning paradigm wholly unchallenged. The study is expected to be completed before planning begins and to furnish the information needed to produce accurate time-series forecasts where historical data is lacking. Intellectually, market studies neatly separate the acquisition of knowledge from its exploitation. Yet, like many good-sounding ideas, it fails in practice.
We return to this link in Chapter Decisions, where we show how decisions can—and often should—be tuned to harvest the value of reduced epistemic uncertainty.
High-dimensional forecasts
Because “forecast” is almost always taken to mean “time-series forecast”, other forms are overlooked. The time-series formalism is restrictive and covers only a fraction of real-world supply-chain situations. Yet only a lack of imagination keeps us from exploring broader kinds of forecasts.
In practice, a time-series forecast is one-dimensional. Even when the horizon spans many periods, each date’s error is evaluated in isolation; the multi-period forecast collapses into a stack of independent one-period forecasts. Whether the numbers come from one shared model or unrelated formulas is immaterial to evaluation: standard accuracy metrics treat them as separate guesses and ignore how errors co-move across the horizon. That blindness matters as soon as decisions couple dates, items, or locations.
To show why high-dimensional accuracy matters, consider an idealized two-product case. Consider a store selling two distinct products that are a perfect pair of substitutes130. The products expire at the end of the day, and are replenished daily. Demand for the two products is 10 units per day; customers pick randomly between them while stock remains.
Three candidate daily forecasts illustrate the point:
-
(A) 5 units and 5 units
-
(B) 10 units and 0 units
-
(C) 6 units and 6 units
Given substitution, a suitable accuracy metric must assign the same loss or value to (A) and (B). Indeed, both anticipations are equally good—indeed perfect under our simplistic assumptions. However, the metric must score (C) worse, since it overestimates total demand by 2 units. No one-dimensional metric can satisfy both requirements. Only a measure that looks at the two items jointly recognizes that (A) and (B) tie while (C) is worse. All classic scores (mean absolute error, mean square error, and so on) wrongly rank (C) above (B). The only way forward is to broaden the perspective to a 2-dimensional accuracy metric.
The example exposes the limits of a one-dimensional view; to see why they matter, we now turn to a more realistic, probabilistic example.
Consider a grocery store and its assortment problem: the choice of the exact list of distinct articles to put on the shelves. Customers rarely buy a single item; most leave with a basket of several products. A superior assortment yields bigger or more profitable baskets on average.
Every assortment still faces stockouts: when a shelf is empty, customers may pick a substitute. Substitution mitigates the stockout. Conversely, the customer may forgo not only the missing article but also the complementary articles he intended to buy. Contagion worsens the stockout.
Only by examining historical market-basket data can we study these substitutions and contagions. Indeed, when market-basket data are available, we have a natural entry point to assess how products relate to one another from the customer’s viewpoint. Furthermore, if loyalty cards are used, sequences of market baskets — each attributed to a distinct customer — can be reconstructed. Loyalty data supplements market-basket data for such analysis.
The technical details are not needed here. What matters is that the relevant information lies in market-basket data. Without it, no matter the sophistication of our endeavor, the behavioral analysis fails for lack of it.
In contrast, when sales data are flattened into time-series, we do not even know how many distinct customers walked into the store on any given day. We cannot see substitutions and contagions because all that is left are tentative correlations between time-series that are both sparse and erratic. Time-series alone cannot support a behavioral analysis.
Circling back to the assortment challenge, note that a market-basket forecast — conditioned on the chosen assortment — offers a straightforward solution. If we can predict future market baskets under a revised assortment, we can assess whether it will be superior to the current one — at least in terms of sales volume.
Interpreted as a point forecast, a “reliable” market-basket forecast is a category error. A single basket often mixes ten or more SKUs; combinations explode, and shoppers freely substitute at the shelf. Expecting to name the exact basket in advance seems—and is—impossible; hence the idea is routinely dismissed. The mistake is the framing: what matters is not one guessed basket but the probabilities of co-occurrence and the lift items exert on one another—precisely the structure we turn to next.
Yet, viewed probabilistically, the outlook is starkly different. Clearly, not every conceivable market basket is equally likely. For example, any basket containing a pair of products never observed together in history is improbable. Establishing a probability distribution over the market-basket space may look technically daunting, but the machine-learning community has developed a sizable “bag of tricks” for such predictions.
The point-forecast perspective, by flattening everything around the “one true future”, gives the misguided impression that the quality of an anticipation can be assessed pointwise, i.e., one dimension at a time; yet this ignores the dependencies within the system. Under a probabilistic perspective, those dependencies become evident through the implausibility of certain outcomes. Demand can rise or fall, but it is implausible for it to rise in odd-numbered postal-code areas while falling in the even-numbered ones. This observation implies a high-dimensional perspective.
Within supply chains, such dependencies are ubiquitous, precisely because—as we have seen—they are systems. One product’s success can cannibalize its substitutes; a grounded jet can cascade into missed connections; cheaper energy can lift an entire sector’s margins. The opening of a flagship superstore may attract new patrons who marginally improve the business of all nearby stores, whatever they sell. Similarly, a successful marketing campaign can lift the sales of all the products of a given brand.
Designing high-dimensional accuracy metrics is technical and lies outside this chapter. What matters is the principle: joint forecasts must be judged in the geometry of the decision—as a whole, not coordinate by coordinate. Proper scoring rules exist for multivariate distributions and let us compare models on a simple criterion: do they concentrate probability where reality tends to fall, while preserving the dependencies that drive economics? With that foundation in place, the remaining gap is agency: a forecast that ignores our own levers still projects the status quo. The cure is to condition predictions on the policy the firm intends to apply. We call such policy-conditioned anticipations functional forecasts—the subject of the next section.
Functional forecasts
The future state of the supply chain depends on decisions yet to be made. Decisions by third parties—customers, suppliers, competitors—should be forecast like any other uncertainty. Indeed, those decision-making processes may well forever remain black boxes for the company. By contrast, the company’s own forthcoming decisions warrant different treatment. After all, the company needn’t guess its own decisions; it controls its decision-making.
As commonly used, forecasts grant the firm little agency. They describe what is likely under frozen policies, not what could happen under alternatives. They condition on decisions already taken when the baseline is computed and remain blind to those still on the table. A manufacturer may cut prices today and, quite properly, see forecast demand lift; yet the same forecast cannot speak to next month’s price move or a promotion not yet chosen. In short, the baseline projects the status quo—a “do-nothing-new” counterfactual—useful for orientation, not a substitute for deciding.
To see why unfinalized decisions matter, consider a fashion brand. The brand runs two collections per year, with products ordered ahead of the season. Assume no in-season replenishments. The brand displays products at the season’s start at full price. For each product, planned quantity matches expected seasonal sales, with a stockout at season’s end. Quantities are set so no end-of-season discount is expected.
Yet by season’s end, some products underperform expectations, leaving unsold inventory. The company’s pricing policy dictates steep end-of-season discounts on underperformers, precisely to liquidate stock.
Under dynamic pricing, the brand can rationally open with larger stocks than under flat pricing. If demand proves higher than forecast, the extra inventory yields profit. Conversely, if demand is overestimated, the policy still liquidates inventory with minimal loss. Pricing is no technicality; it is integral to the brand’s flow strategy.
A “regular” demand forecast, probabilistic or not, cannot capture the sales trajectories induced by this policy. When the baseline forecast is computed, there is no reason for projected demand to exhibit—at season’s end—a late spike from discounts. Discounts arise only when demand was overestimated; with accurate demand, initial stock would be lower and no end-of-season discount would be needed. The policy makes demand partly self-correcting: sales align with stock levels only insofar as discounts are aggressive.
A functional forecast, by contrast, explicitly incorporates one or more decision rules. Those policies are taken as given; they are not learned from historical data. Moreover, those policies remain isolated within the forecasting logic. Thus, for a model to be “functional” (policy-conditioned), one should be able to swap a policy for an alternative and directly obtain the revised forecast. In that sense, a functional model is incomplete: without the required policies, it cannot produce forecasts.
As we will see next, functional models need not be much harder to implement than their unconditioned counterparts. However, assessing their validity is a thorny challenge. Indeed, as long as the policies fed to the functional model remain close to those historically practiced by the company, a well-engineered model should behave sensibly. However, if a policy ventures into untested territory, the model’s validity becomes uncertain.
Revisiting the case above: if the brand historically used end-of-season discounts between 15% and 30%, the data support modeling any combination within that range. But if the brand runs its functional demand model with discounts between 30% and 60%, any figures produced will be, at best, highly speculative. Those predictions lack direct empirical support; extrapolating customer response to untested prices may prove very surprising.
Unfortunately, beyond narrow situations, judging whether a functional model will return sensible figures under an arbitrarily novel policy is a problem of general intelligence. Assessment must be tailored to the specifics. For example, trial the new policy on a limited scale solely to acquire further empirical data. This process is exploration, discussed in the next chapter.
Without functional forecasts, a company merely projects the status quo, replicating its past policies. If the company wants to evolve, it needs the capacity to project how envisioned policy changes will affect the flow.
At small scale, functional forecasts are obvious, though seldom named. Consider a neighborhood baker choosing his opening hours. Longer hours enable more purchases—early commuters and late shoppers can buy—yet consume a scarce resource: time with his family and rest. The relevant forecast is not a single number (“tomorrow’s demand is 180 loaves”) but a function linking decision to outcome (“expected sales as a function of opening hours”). By selecting the input—the schedule—the baker partly manufactures the demand he intends to serve; this is the functional view.
At scale, when thousands of forecasts are generated, many firms lapse into status quo projections driven by bureaucratic inertia. Naturally, this attitude serves them poorly. Unfortunately, most supply chain textbooks—and the software built on them—embrace the same perspective.
Technically, functional forecasts matter for two reasons.
First, a software-engineering reason: they enforce a clean separation between the predictive core (models that turn records into probabilistic statements about demand, lead times, prices, returns, and costs) and the reified policy (explicit code and parameters that implement pricing, allocation, service rules, and other levers). Once disentangled, the same predictive core can be exercised under multiple admissible policies; policies can be swapped without retraining; and the economic effect of each rule becomes auditable.
Second, a statistical reason: predictors learn only what past policies exposed. When a policy ventures outside that data support—new price ranges, promotion cadence, service promises—the forecast becomes an extrapolation. Under such a policy shift, precision degrades and uncertainty widens; results should be treated as provisional until measurements accumulate under the new policy.
With these constraints in view, functional forecasting builds the bridge from “what is likely” to “what we would do if …”. The natural instrument to exercise that bridge at scale is a simulator—a stochastic, discrete model that embeds policy and propagates uncertainties. We now turn to it.
Stochastic simulations
A simulation is a model that imitates a real-world system and lets us observe how its state evolves over time. As computing power grew, simulators proliferated, including supply-chain versions often marketed as “digital twins”131. Properly understood, a simulator is a high-dimensional probabilistic forecast: it produces many plausible futures—with implicit frequencies—rather than one ordained trajectory. Used this way, simulation is a legitimate instrument for anticipating future supply-chain states. Yet because this probabilistic nature is rarely made explicit, outputs are too often taken at face value; assumptions go unexamined, and few bother to check that the simulated distributions resemble what operations actually deliver.
Supply-chain simulators vary widely, yet nearly all share three traits: they are microanalytic, discrete, and stochastic.
A microanalytic (bottom-up) simulator decomposes the system into minimal units governed by simple rules. Units include agents—customers or suppliers—and inanimate entities such as SKUs. The microanalytic view assumes system complexity is merely apparent and arises from repeated application of simple rules.
A discrete simulator (discrete-time or discrete-event) proceeds through a sequence of events. Each event occurs at a particular instant and marks a change of state in the system. Between events, nothing changes—an expedient that vastly simplifies the simulator’s software implementation.
A stochastic simulator (often called a Monte Carlo simulator) uses non-deterministic rules that produce random outcomes. Thus, repeated runs from identical initial conditions produce different results. Use deterministic rules when possible; reserve stochastic rules for unknowns, reflecting our imperfect knowledge of the system’s evolution.
Simulators are instruments of functional forecasting: they map explicit flow policies to predicted outcomes. By encoding decision rules—what to buy, make, move, and price—and when—along with operational constraints, the simulator produces forecasts conditioned on those rules. That conditioning is the point: it puts alternative policies on equal footing and lets us quantify their economic consequences.
Bypassing that conditioning almost always forces crude approximation. For example, treating demand as unaffected by quality of service—assuming repeated stockouts never depress observed sales—blinds the simulator to the very behavior the policy induces. Such a model systematically overstates attainable volumes, misprices options, and steers decisions off course. Hereafter, “functional forecast” denotes this policy-conditioned prediction rather than an unconditional time series.
The software implementation of such simulators is conceptually straightforward. A main loop traverses the edges of a graph representing the flow. Nodes typically represent SKUs—or aggregates when finer granularity is too costly. Each node carries a small set of rules, either deterministic—e.g., decrement stock per outgoing unit—or stochastic—e.g., draw random demand for the period. Each iteration of the main loop represents a discrete time step—typically a day, possibly shorter.
If one doesn’t look too hard at validity, such a flow simulator can often be built in a few days. Hence its popularity with enterprise software vendors. The functional design is appealing: it yields arbitrarily fine-grained policy predictions. Yet without rigorous validation, that trust is misplaced.
In fact, simulator outputs can be—and often are—outright nonsense. A rule’s plausibility does not guarantee its correctness, nor does pairing two valid rules ensure the simulator combines them correctly. Mundane coding errors also creep into the simulator’s codebase. As noted earlier, novel policies may undermine the stochastic rules if the simulation enters “uncharted” territory. Ultimately, validity is empirical and must be settled by measurement.
A duality links probabilistic forecasts and stochastic simulations. From a probabilistic forecast, it is possible to generate samples according to the underlying probability distribution (such samples are usually referred to as deviates). From a complete probabilistic forecast, one can derive the stochastic rules that drive the system; conversely, observing a simulator reveals the probabilities of all outcomes. Thus a simulator can estimate the complete probability distribution over all admissible states; hence a probabilistic forecast can be derived from a simulator.
In practice, numerous technicalities hinder conversion between a probabilistic forecast and a simulator. The probabilistic forecast might not be sufficiently fine-grained to specify all the simulator’s rules. Conversely, the number of runs needed to estimate all the probabilities may be prohibitively large. Even so, the duality implies that stochastic simulations share the accuracy metrics of probabilistic forecasts.
Therefore, treat a simulator like any forecast: quantify its accuracy. The work is technical; ignoring quantitative assessment invites costly mistakes.
Decisions
Supply-chain practice is the continual allocation of scarce resources under uncertainty. The aim is not a plan but a profitable flow: convert cash into goods, goods into service, service back into cash—while preserving the options that future information will render valuable. Even mid-sized firms make thousands of such allocations daily—and the largest, millions—so the decision system cannot be ad hoc. It must make choices explicit, price them, and then sequence and execute them so that unattended software can issue them safely at scale. Systematization does more than guard against a rogue buyer exhausting working capital; its purpose is to produce decisions that, in expectation, earn superior risk-adjusted returns.
To close the gap between principle and practice, we now sharpen the notion of a decision first sketched in Chapter 1.
A definition, however crisp, leaves two practical questions. First, where does a decision begin and end once time and randomness enter? Second, by what single criterion are alternatives compared when they draw on the same scarce pools—cash, capacity, attention? In supply-chain settings, a decision is not a line on a plan. It is a flow commitment taken at a granularity fine enough to change economics, made under uncertainty, and recorded in a ledger that future events will audit.
Rendering such commitments computable and auditable requires bringing four ingredients to the surface. The frame comes first: the present sphere of agency—admissible moves, feasibility tests, cadence, and the few ratchets that erase tomorrow’s options—must be written—ideally as code. Next comes a probabilistic view of near futures: demand, lead times, returns, reliabilities, and any other distributions that materially affect outcomes. Third, all penalties and rewards must be expressed in money and attached to owners—stockout pain, obsolescence, capital charge per unit time, congestion, late fees, the option value of waiting—so unlike frictions become commensurable on one ledger. Finally, marginal commitments are ranked by their expected, risk-adjusted aperiodic rate of return; those that clear the hurdle are issued, logged, and later judged against realized outcomes.
The chapters so far have prepared this stance: optionality and variability (Chapter 1) supply the space of moves and explain why a probabilistic lens is mandatory; economics (Chapter 4) provides the coin-denominated objective; information and intelligence (Chapters 5–6) delimit the roles of records, reports, and decision engines. The present chapter turns these pieces into practice. It begins by thinking the decision: surfacing the frame, banishing artifacts, opening the economics to inspection, and restoring reasoning at the margin. Subsequent sections examine unattended execution and the optimization machinery suited to repeated, uncertainty-laden choices, so that each commitment becomes a small, priced wager the firm can both defend ex ante and audit ex post.
Thinking the decision
A decision, as defined above, is a commitment that erases alternatives. Thinking the decision therefore begins one step earlier, with the act that selects which alternatives are even seen and how they will be priced. In most firms, this prior act remains tacit. People talk in the grammar of familiar “policies”—FIFO, full trucks, minimum order quantities, shelf plans—so the choice appears obvious and the unchosen, invisible. The danger is not clerical error but metaphysical slippage: a policy silently substitutes for the economic question it was meant to answer. Reversing that slippage—bringing the question back to the surface, keeping artifacts in their place, making the economics explicit, and reasoning at the margin—turns decision‑making from ritual into a scientific practice fit for automation.
Surfacing the frame
Managers are tempted to think with solutions—not a folly, but a correct and useful heuristic. The space of conceivable “problems” is boundless—designing an anti‑gravity engine is a problem too—and without the scent of a solution most inquiry would be wasted. Yet policy‑first thinking fossilizes quickly. “Serve stores FIFO” begins as a decent rule of thumb and, over time, hardens into doctrine; each exception is patched by another rule, and soon the firm steers by accreted case law.
What must be surfaced, before any policy, is the frame. It is the concrete, inspectable boundary of agency for the present decision cycle: what can be altered now and what cannot; which options are admissible; which random elements will matter; which commitments would ratchet the future; and at what prices this slice of reality will be judged. The frame should be written, not merely described. In a modern organization, it lives as code: a short program that enumerates admissible moves, encodes the feasibility tests, and calls the valuation function. If the program is long, the frame is muddled; if the program is opaque, the frame is political.
Once framed, the same situation admits different moves. Consider an importer that fills containers “product‑by‑product until full”, deferring the rest to the next sailing. The policy sounds prudent; the frame it hides is narrow indeed. Make the admissible options explicit—co-loading across suppliers, temporary cross-docking to other regions, price nudges synchronized with sailing dates, late binding between look-alike items—and then the optimization no longer searches within one corridor but roams a small atlas. The gain comes less from a cleverer solver than from a better question. Problem design is not icing; it is the cake.
Banishing artifacts
Once a frame exists, numerical artifacts will appear in the machinery: forecasts, buffers, service levels, ABC tags, budgets, and their innumerable cousins. They are often useful as intermediate representations, just as a chemist uses reagents that never leave the bench. Trouble begins when an artifact escapes the lab and is enthroned as a target. Goodhart’s law then takes over: the measure becomes the goal and ceases to measure.
Service level is the classic escapee. It began life as a proxy for a money‑denominated trade‑off between shortage pain and carrying cost; it ended as a percentage that commands the organization. Planners raise it to feel safe, finance lowers it to free cash; neither side is forced to state prices. Forecast accuracy plays the same role for teleological planning: its improvement is celebrated even when profit falls.
Artifacts should be kept on a leash. Inside the decision engine, they are acceptable as transient coordinates that make a computation tractable; outside, they have no standing of their own. A replenishment is not “good” because its SKU hit 97% service; it is good because, given the options, constraints, and probabilistic futures considered in the frame, it maximized the firm’s risk-adjusted rate of return. The engine may well compute a distribution over daily demand, a shadow “service level” implied by the shortage penalty, and a working‑capital charge per unit per day; it may even print them for inspection. But dashboards and incentives must anchor on decisions and their realized economics, not on the laboratory glassware that helped to compute them.
There is a second reason to banish artifacts from the altar: they are systematically harder to value than the decision itself is. A safety stock number has no invariant economic meaning across items, seasons, or geographies; its virtue is parasitic on the hidden penalties and opportunities it stands in for. Turning such a chameleon into a KPI invites bureaucracy to colonize the flow. The cure is to treat every artifact as a private, provisional object, recomputed as needed from authoritative records, never hand‑maintained, never made a target, and discarded as soon as a better internal representation presents itself.
Artifacts are not only numerical. Organizations also mint bureaucratic artifacts—policies and partitions—that began as rough safeguards for attended processes and later ossified into rules that substitute themselves for the economic question. When such artifacts leave the lab and acquire institutional standing, they distort decisions exactly as service levels did: the means becomes the end.
As British historian Cyril Northcote Parkinson cautioned in Parkinson’s Law (1957), “work expands so as to fill the time available for its completion.” Bureaucracies grow by reflex; large organizations running sizable supply chains are no exception. As they swell, extraneous rules and internal partitions accumulate in their decision-making processes, surfacing mainly as made-up constraints and made-up silos—both of which sap efficiency and blur accountability.
Made‑up constraints typically arise after a mishap. Roger from purchasing orders eighteen months’ worth of sales by mistake; in response, management imposes a policy capping purchase orders at “six months of sales”, measured by a moving average. In manual settings, clerks enforce this cap with common sense: they ignore it for new products with no history, for highly seasonal items whose comparable is twelve months back, for fast‑growers whose future dwarfs the past, or for planned promotions whose lift is engineered. An unattended engine cannot “sensibly ignore” rules; it either obeys the letter or is declared noncompliant.132
The remedy mirrors the treatment of numerical artifacts: keep such rules private to the engine and express them as feasibility tests and priced penalties, not as public commandments. If buying above a naive “six‑month” cap is occasionally rational, the cap belongs in the objective as a coin‑denominated nuisance, not as a veto; newness, seasonality, and engineered lifts become explicit features of the frame, not tribal exemptions. A rule written as code, priced and owned, is legible and contestable; a rule written as a decree invites gaming and stifles automation.
Made-up silos fracture a single economic question into a procession of artifacts so that lightly tooled teams can cope. A fashion retailer, for instance, runs the warehouse-to-store flow through three artifacts: an assortment list (what a store may carry), a launch pack (how much at start), and a replenishment rule (what to send later). The choreography looks tidy; it quietly manufactures constraints. The assortment list pre-commits shelf real estate and vetoes substitutions that early sell-through would justify; the launch pack pre-allocates labor, line-haul, and facing capacity that replenishment must then buy back—often at a premium; the replenishment rule then pretends to optimize what is left. Money is left on the table because the underlying question—where does one more unit earn the highest risk-adjusted return, given current constraints?—was broken into artifact-bounded procedures before it could be asked end-to-end.
Here again, artifacts must stay on a leash. Silos may curate those artifacts; they must not own the decision. The engine should see the end‑to‑end frame—admissible moves, shared constraints, and prices—and produce a single flow of commitments. If governance requires organizational boundaries, the interface must carry prices and feasibility, not quotas and vetoes: the distribution center (DC) publishes shadow prices for scarce resources; stores publish shortage penalties; assortment is a proposal the engine may expand when economics justify it. In other words, keep the ceremony parochial and the economics unified.
Once artifacts—numerical and bureaucratic—are domesticated as private, priced, and provisional objects, the stage is set for the only debate that matters: the prices themselves. The next section turns to that point directly.
System economics
The previous chapter established money as the firm’s common scale. System economics requires that every decision be priced from the vantage of the whole firm, not a departmental KPI. In practice, this means the prices inside the objective function—stockout pain, obsolescence, shelf congestion, late fees, working-capital cost, option value of waiting, goodwill loss—are stated once company-wide, attached to named owners, versioned, and exposed. Concealed prices do not become neutral; they bias the flow toward whoever smuggled them in. A hidden penalty for backorders is a private agenda masquerading as science.
Making the economics system-wide does not require a plebiscite for every coin; it requires numbers legible and contestable across functions. Finance should publish the firm-wide cost of capital and liquidity limits; operating teams should translate them into per-unit, per-day charges reflecting inventory half-lives and salvage values. Merchandising should not demand “98 % availability” without pricing the next point in coins per day; logistics should not insist on full-truck thresholds without pricing the congestion and spillover they avert. Where markets exist, borrow external prices; where they do not, mint private valuations and defend them. To arbitrate across silos, scarce shared resources—dock slots, shelf space, line time, capital—should carry shadow prices: engine-computed, provisional numbers expressing the current opportunity cost of using one more unit. They replace quotas and vetoes; they are not a new policy. We return to them in the discussion of sequential decisions below. The goal is not unanimity but comparability on the company P&L. Once penalties and rewards are expressed in the common coin, options that used to live in different provinces—“grant a refund beyond policy”, “buy one more unit of a long-tail SKU”, “hold a truck to top up a route”, “extend the return window for this class of customers”—compete on the same ledger.
This system view also clarifies accountability. If a decision systematically underperforms once outcomes materialize, the dossier shows whether the failure lies in the frame (options omitted), the forecasts (poorly calibrated tails), or the prices (penalties misstated). Corrections then target the offending ingredient rather than diffusing into vague calls for “better collaboration”. Local KPIs lose their alibi: a move that flatters a departmental metric but destroys coins elsewhere will be exposed by the aggregate ledger. Hidden prices breed politics; open, system-level prices breed performance. With company-wide prices in place, the engine returns to its natural stance: reasoning at the margin under a single ledger.
Embracing the margin
Economics is a marginalist discipline; so is supply chain when practiced properly. The relevant question is almost never “what is the right stock level for this SKU this season?” but “what is the risk-adjusted return of one more unit—here, now—given everything already committed?” The answer is a number in coins per unit per unit time, and it shifts as the commitment shifts. Thinking at the margin forces grain, time, and contingency into view.
Grain first. Decisions should be computed at the finest granularity that materially changes their economics. For commodities, the unit is a piece; for transport, a stop or a cubic meter; for aviation rotables, a serial number whose cycles and hours render two parts with the same P/N economically incomparable. An airline can rightly sell a given unit to a rival’s AOG (Aircraft On Ground) request while keeping another ostensibly identical one to hedge its own tail risk; the marginal values diverge because the insured futures differ.
Time next. Marginal values are aperiodic: they must account for the half-life of the load, not just the lead time. A crate that clears in two days with deferred payment can be quite profitable, though the same crate, paid up front and sold over eight weeks, is mediocre. When the engine ranks admissible moves by their aperiodic rate of return—as argued in Chapter Economics—it naturally prefers quick-cycling bets and penalizes long ratchets unless their upside is commensurate.
Finally, contingency. Margins live on distributions, not on points. The extra unit that looks profitable under the median forecast may be ruinous once a fat-tailed lead time is acknowledged; conversely, a unit that seems wasteful under a thin-tailed model becomes attractive when a rare but catastrophic stockout is correctly priced. This is why a probabilistic forecast is not a luxury: it is the substrate on which the marginal calculus operates.
At scale, embracing the margin turns the firm’s daily work into a continual resorting of millions of micro‑options. Each new commitment reshapes the frontier: shadow prices emerge on tight constraints, rival moves change the payoffs, yesterday’s best next step becomes today’s third‑best. A static plan cannot price moving edges; a decision engine can. It does not conjure certainty; it computes, at each instant, the best next wager and writes it into the ledger with reasons. The rest is merely bookkeeping.
Thinking about the decision this way—frame explicit, artifacts domesticated, economics open, margins sovereign—does not add ceremony. It subtracts superstition. It also aligns with the rugged vision of the previous chapter: decisions become small, auditable wagers placed under uncertainty, guided by prices the firm is willing to defend, and revised as reality reveals itself.
Why automate
Computers are unreliable, but humans are even more unreliable. Any system which depends on human reliability is unreliable. One of the many Murphy’s Laws.
The previous pages defined a decision as an auditable commitment that erases alternatives and is justified only by its expected risk‑adjusted rate of return. Once the frame is made explicit and the economics opened, a supply chain becomes what it has always been in substance: a mill of innumerable small wagers whose edge is earned at the margin. At modern scale, the relevant edge is inaccessible to unaided clerks. What exhausts them is not arithmetic—spreadsheets already shift that burden—but the sheer combinatorics of admissible options, the demands of probabilistic reasoning under fat tails, and the continual repricing of constraints as reality moves. If the bottleneck is felt in SQL dashboards rather than in steel and diesel, the firm has chosen to limit itself in the one realm where limits are least binding—computation.
There is a second, quieter reason to automate: repetition under control. A manual or semi-manual process changes every time people change; staff turnover, local fixes, and shifting attention make yesterday’s baseline incomparable with today’s. By contrast, an unattended process can be duplicated, perturbed, and measured. Two numerical recipes can run concurrently against the same ledgers—one in authority, one in shadow—while drawing on the shared staff because neither demands daily supervision. Measured this way, progress ceases to be a matter of opinion; it becomes an audited delta in coins.
Spreadsheets made clerical work tolerable; they never made it effective. Their “logic” is a tangle of ad hoc formulas that neither encode a frame nor price uncertainty. The result is familiar: last-minute overrides, safety buffers stacked on safety buffers, and a ritualized haste in which a clerk must finalize a hundred quantities before lunch. It is tempting to imagine that granting everyone more time would yield better answers. In practice, the pause inflates headcount, deepens silos, and delays commitments—each delay an opportunity cost that rarely appears on the dashboard but always appears on the P&L.
Automation, in the present sense, is not “using more software”. It is the institutional decision to let systems of intelligence—introduced earlier—compute, issue, and log decisions unattended under normal conditions. What follows examines this stance from several angles: as an asset, as a discipline for overrides, as a cultural turning point, and as the natural outcome of a century‑long trajectory from management to machinery.
A productive asset
Automation turns decision‑making from perishable labor into a capital good. A system of intelligence is not a report that “assists” a planner; it is an engine that commits coins on the firm’s behalf according to an objective it can defend. The expense profile flips: what was OPEX—hours spent nudging cells—becomes CAPEX—a codebase whose operating cost is negligible relative to the volume of choices it processes. Each engineering day spent improving this asset is accretive: either decisions get better, compute gets cheaper, or upkeep gets lighter. If none of these occurs, no further money should be spent.
This asset is productive because it encodes the frame and the prices. Unstated heuristics that once lived in memos and meetings migrate into version‑controlled code. The engine reads authoritative records, infers the distributions that matter (demand, lead times, returns, repair times), evaluates admissible options under those distributions, and writes back auditable commitments. It does so at the cadence reality demands, not at the pace a committee can endure. Dual-run becomes routine: the incumbent recipe holds the pen while a challenger produces shadow decisions, and a clean A/B delta accumulates in the ledger. When confidence drops—the tails look miscalibrated, a constraint begins to bind unexpectedly—halting heuristics suspend emission; the fallback is not a thousand exceptions but a stop and a fix.
Software is cheap to run and, if kept small and legible, cheap to maintain. Its performance limit is not human patience but the economics encoded in it. Any genuine insight—about substitution, deferring or expediting, rerouting, the option value of waiting—can be folded into code and thereafter applied at machine scale. The clerical ceiling disappears; what remains is the strategic ceiling set by the firm’s ability to price what it values.
Manual overrides
Automation always raises the question of when to intervene. An emergency stop belongs in any powered machine; routine “manual corrections” usually signal a design flaw.
Vendors have long softened brittle automation by deputizing users as human coprocessors. Alerts and “exceptions” flood inboxes; planners are asked to inspect most outputs and override many. This pattern defeats the purpose. If a system needs a person to catch bad decisions, that person will soon distrust it, and the organization will revert to spreadsheet theatre with extra steps.
The cure is architectural rather than hortatory. First, the engine must be built so that every decision it emits is sound on its own terms: framed within admissible options, priced in coins, and justified under a probabilistic view of the future. Second, when unsure, the engine must halt—not pester. Halting is not abdication; it buys time for the only interventions that durably improve outcomes: edits to the frame, the prices, or the forecast semantics. Third, trust is earned by design: dual-run before cutover; visible dossiers for every commitment; explicit, versioned valuations; and measured deltas after the fact. Chapter Engineering will call this discipline experimental optimization.
Overrides do not vanish; they change nature. A rare, targeted override edits a valuation (“the shortage penalty for this class of customers was understated”), narrows the option set (“do not ship lithium batteries by air this month”), or repairs a semantic fault (“this field changed meaning after the ERP patch”). It does not scribble a different number on a purchase order while leaving the machine untouched.
Culture and unrest
Rising productivity invariably unsettles incumbents. Tiberius refused cranes; the Luddites smashed looms. Modern firms are spared neither anxiety nor politics, and white-collar workers face a novel twist: they are asked to help automate their own jobs.
The paradox dissolves once roles are seen clearly. The work to be mechanized is the clerical grind of enacting yesterday’s heuristics. The work to be elevated is the articulation of frames and prices, the stewardship of semantics, and the rapid rewiring of logic when the world shifts. Software cultures solved this tension decades ago. They prize people who make themselves obsolete at the keyboard—because value moves to the next problem: the tool becomes the asset; the engineer moves on. Where this culture is resisted, markets—not memos—adjudicate. Firms that refuse to compound the asset are displaced by those that do.
Executives cannot purchase this culture off-the-shelf, but they can remove the incentives that poison it. Reward decision quality, not seat counts. Pay vendors for outcomes, not logins. Promote the contributors of better frames and valuations, not the curators of larger dashboards. Above all, give the few who can actually rewire the engine the mandate to do so quickly—their leverage dwarfs a roomful of human schedulers.
Two recurrent patterns merit daylight. First, headcount is still treated as a proxy for status. Managers who grow the number of “seats” under them gain budget and influence; automation, which collapses seats, feels like a demotion. Reverse the currency. Tie prestige to uplift per person and per watt. Make shrinking attended seats an explicit objective and celebrate the teams that make themselves obsolete at the keyboard; in a software culture, this is advancement, not abdication.
Second, per‑seat licensing and “engagement” metrics make vendors complicit in keeping humans in the loop. A system that bills by logins, clicks, or “active users” is economically opposed to unattended flow. Contracts should be priced on outcomes the firm wants—volume of unattended commitments issued with audit trails, measured uplift against a dual-run baseline, time to rewire logic after a frame change—not on the clerical motion the tool induces. When money is paid for results rather than for screens, incentives align with automation rather than against it.
Finally, beware of metric theater: when bonuses hinge on accuracy percentages or on-time scores detached from coin-denominated valuations, energy flows to improving the indicator rather than the economics.
Human exceptionalism
The claim that “a computer must never make a management decision because it cannot be held accountable”133 seems humane; it is, in fact, confused. Most large organizations already run vast tracts of policy that leave no room for discretion. A credit hold on overdue accounts is enforced whether a person clicks the button or a program does. Accountability attaches to the policy’s authors and the executives who endorsed the prices and constraints it encodes, not to the clerk who executes it.
Insisting on a human hand for symbolism degrades both speed and quality. It also invites a worse failure mode: invisible responsibility. When a rule lives in a spreadsheet and a routine exception is scribbled into a cell, no one learns. When the same rule lives in code and an outlier forces a halt, the design is amended in daylight. The relevant test is economic, not anthropomorphic: given the same inputs, the same admissible options, and the same audit trail, does the system yield higher risk‑adjusted operating profit than the incumbent? If yes, it is “intelligent enough” for the task.
From management to machinery
The puzzle is not why automation is desirable, but why it took so long to become routine outside a few niches. For most of the 20^th^ century, decision processes were necessarily managerial; only human minds could digest information and emit sensible choices. The question then became how to divide the labor across many minds. Consulting authors proposed elaborate fabrics of committees and calendars to make the division bearable. The fundamental flaw was silent: what is “best” for a fragment seldom aligns with what is best for the whole. A homeware division that deploys one million coins at an expected return of ten percent can, without malice, preempt the kitchenware division from deploying the same million at fifteen percent. However finely the organization is sliced, all managers draw on the same finite pools of cash, inventory, capacity, and attention.
Bureaucratic methods tried to heal this with more communication. Meetings, memos, and escalations transported fragments of knowledge upward, then downward, hoping that alignment would follow. Yet allocation is quantitative; without a common ledger of prices, coordination reverts to hierarchy. Operations research, born in wartime, promised a different future: let algorithms optimize the allocation end to end. Hardware constraints, crude data semantics, and a taste for deterministic toys conspired to blunt that promise. By the late 1970s, Russell Ackoff was already writing its obituary.
Spreadsheets filled the vacuum. They multiplied clerical capacity but froze decision‑making at a pre‑scientific stage. Attempts to replace them with “planning” modules failed because those modules inherited the same deterministic posture and the same artifact worship. Users drifted back to files; vendors added more screens. The decision load returned to white-collar ranks, now larger than the blue-collar ranks that move cartons.
The economics never changed: repetitive supply‑chain decisions can and should be automated. The enablers matured quietly. Systems of records stabilized; probabilistic methods, once exotic, became pedestrian; hardware became cheap; and a new architectural species—systems of intelligence—emerged to sit alongside ledgers and reports rather than pretending to replace them. Fully unattended decision engines have been in production for more than a decade in first-movers134. The pattern will spread, and spreadsheets will keep their place as Swiss-Army knives—handy in a pinch, absent from the production line.
Two further obstacles, often overlooked, explain the long delay. First, distributing a decision across many heads is intrinsically treacherous. No natural ordering exists that makes local choices commute. Advertising budgets, shelf space, and factory capacity for a new product constrain one another; whatever team “goes first” traps the others into an inferior corridor. The correct resolution is not another committee, but a unified optimization instance that draws on each team’s expertise where it belongs—in valuations, constraints, and scenarios—while letting the solver synthesize a single commitment. Yesterday’s commitments, once enacted, re‑enter as data; the next cycle proceeds on a narrower but clearer stage. Second, organizations accumulate made‑up constraints to protect themselves from their own past errors. Manual processes survive by ignoring these rules when common sense demands it; software cannot. The cost of automation is therefore paid up-front in the removal or re-pricing of such constraints; the dividend is paid forever in cleaner frames and higher returns.
The broader arc is familiar. Over the last three centuries, profit‑seeking entrepreneurs relentlessly substituted capital for labor, and blue‑collar shares fell accordingly. In many industrialized regions, the headcount managing the flow now exceeds the headcount physically moving it135. The next increment of productivity will come from automating away the clerical part of white‑collar work. Resistance will be spirited; the economics will be patient—and decisive.
Finally, a word on the technological “why not earlier”. Until recently, three preconditions were missing. Ledgers were too fragile to serve as authoritative ground truth; probabilistic reasoning was too cumbersome to be embedded in everyday code; and vendors were rewarded for selling CRUD at intelligence prices. All three have shifted. Systems of records have plateaued in value and reliability; probabilistic toolchains fit on a laptop; and buyers have learned to separate bookkeeping, reporting, and decision engines. Once these are in place, the case for automation is not rhetorical; it is arithmetical.
In short, we automate not to polish a plan but to restore economics to the driver’s seat. Unattended decision engines put the bottleneck where it belongs—in atoms and hours—while delivering the one precondition for sustained improvement: comparable experiments that run daily, at scale, with coins on the line.
Where to start
Automation is not a slogan; it is an allocation problem: where should the next coin of engineering effort be spent so that unattended decisions repay fastest? In practice, the operative notion is the difficulty of a decision process, understood economically as the inverse of the expected return on investments in improving it. High expected return, low difficulty; low expected return, high difficulty. Two regimes matter most. The base difficulty is the cost of migrating a flow from attended spreadsheets to an unattended decision engine under an explicit frame. The marginal difficulty is the cost of extracting the next increment of performance from a flow that is already automated.
The point is simple and vivid: scale lowers base difficulty. When millions of small, similar commitments recur and the frame can be written down—admissible options, feasibility tests, and money-denominated prices—the first automation passes typically repay quickly. By contrast, small, idiosyncratic niches combine low payoff with high marginal difficulty; they are the worst places to start and should be tackled last, however tempting their apparent tractability.
Consider a grocer whose national distribution center feeds 800 stores nightly. Automating DC‑to‑store dispatch has a short, legible frame: admissible moves are “ship 0/1/2 units of SKU $$i$$ to store $$j$$ tonight”; feasibility tests are truck cube, door slots, and store caps; prices are shortage pain, aging, line‑haul and handling, plus a shadow price on distribution center stock; distributions are next‑day sell‑through and late‑truck risk. One numerical recipe then covers tens of thousands of homogeneous micro‑decisions per day, so a first pass—greedy accumulation under these prices, with a halt when confidence drops—repays quickly. By contrast, starting with a boutique flow—quarterly pop‑up assortments for twelve flagships entangled with visual‑merchandising rules, ad hoc launch events, and planogram politics—offers little scale and murky prices; the “frame” is mostly ceremony. The large river is easier than the puddle: scale makes base difficulty low, while idiosyncrasy makes it high.
Engineers often misread solver timing curves. For any fixed technique, enlarging a model—more variables, more constraints—almost invariably induces superlinear growth in CPU and memory.136 That curve is a property of the method, not of the business problem. Economics supplies levers the curve ignores: coarsen the frame wherever granularity does not move money; shorten horizons to the next moment of control; replace brittle hard limits with priced penalties; and help the solver with warm starts and decompositions so yesterday’s answer speeds today’s search. The goal is not algebraic purity but risk-adjusted return within a sensible compute budget, day after day. In practice, a rough-but-priced formulation that runs reliably outperforms “exact” models that time out—and both beat spreadsheet clericalism once scale is present.
A practical triage follows. Begin where the frame is short and legible: options are enumerable without heroics, feasibility tests are mechanical rather than political, and the prices that govern the objective—stockout pain, obsolescence, capital charge per unit time, the option value of waiting—can be stated, owned, and versioned. Favor flows whose cadence gives the engine time to think (minutes to hours): overseas purchasing and consolidation, DC-to-store dispatch, network stock balancing, repairables provisioning. Defer micro-niches whose economics rest on unresolved, implicit prices—governance problems before engineering ones. Above all, let measurement decide: run challengers in dual-mode against incumbents, attribute deltas in coins within a clear window of responsibility, and promote only when the uplift pays for itself.
Seen this way, “where to start” is no mystery. Attack the biggest rivers of decisions first, accept approximate answers that respect the economics and the frame, and let the unattended engine compound from there. Small problems feel safer; large ones are easier.
Deciding is optimizing
Mechanizing choices requires a formalism that tells a machine what to prefer and when to stop. “Optimization”, a branch of applied mathematics, supplies that formalism—but only once it is recast to fit the economic and iterative character of real decisions. The textbook posture—state a model, hand it to a generic solver, await a provably optimal plan—misfires in practice. What works is humbler and more powerful: an optimization frame that exposes admissible moves, prices their consequences in money, ingests probabilistic views of tomorrow, and is engineered to run—unattended—again and again as the world shifts.
The optimization paradigm
Classical mathematical optimization notation speaks of variables, constraints, and an objective (see the annex). We keep the triad but shift its meaning to match the decision grammar developed earlier in this chapter and in Economics and Intelligence.
Variables are not blunt totals (“the quantity for SKU X this month”) but marginal commitments at the finest grain that changes economics. For some problems the grain is a unit, a stop, or a cubic meter; for rotables it may be a specific serial number whose history makes it economically unique. Thinking in marginals keeps the engine aligned with the rate–of–return calculus: the question is never “what is the right stock level?” but “is one more unit, here and now, worth more than its next best use?”
Constraints are feasibility tests that decide which options may be exercised. They should be written as code that inspects a candidate move and either admits it or rejects it with a reason. In supply-chain reality, very few limits are truly absolute: almost any bottleneck can be relaxed for a price or a delay. It is therefore prudent to move many “constraints” into the objective as priced penalties, keeping only the handful that reflect physical impossibilities or legal prohibitions. This makes unlike frictions—dock congestion, crew overtime, shelf resets—comparable on the same coin ledger and lets the engine find the least-bad compromise under uncertainty.
The objective is a money‑denominated valuation consistent with the firm’s goal: maximize expected, risk‑adjusted rate of return. It is not a percentage target masquerading as economics. A useful objective must be explicit about the prices it assumes: stockout pain, obsolescence, capital charge per unit‑time, late fees, goodwill loss, option value of waiting. Hidden prices breed politics; open prices breed performance.
Forecasts enter as first‑class inputs. Because outcomes hinge on futures that are partly unknowable, the engine does not score plans against a single guessed trajectory but against distributions—over demand, lead times, returns, reliabilities, and sometimes prices. The objective remains a deterministic function of its inputs; the randomness lives in the forecast objects that feed it. This separation keeps causality legible and the blame assignable when a bet underperforms: was the frame narrow, the valuation misguided, or the forecast miscalibrated? The heavy probabilistic machinery is developed in the next section; here it suffices to insist that a decision engine must be built from the start to ingest distributions, not points.
Finally, a word on expression. Mid-century doctrine tried to separate statement from resolution—one “declarative” file for the problem, a black-box solver for the rest137. That separation flatters elegance; it handicaps practice. A decision engine that runs every day benefits when its model helps the method: variable orderings that reflect economics; warm starts from yesterday’s solution; decompositions aligned with geography or lifecycles; caches that remember expensive subcomputations; feasibility tests that fail fast. This is not impurity; it is mechanical sympathy between problem and machine.
On optimality
Within a fixed frame—finite resources, explicit admissible moves, stated prices—there is a best allocation. The point matters: “good enough” is not an abdication of rigor but a recognition that economics, not algebra, decides when a search should stop.
Two lessons follow. First, supply-chain problems exhibit diminishing returns along most levers. Extra inventory improves service until it no longer does; extra expedites tame lateness until they no longer do; extra assortment widens reach until it clogs attention. This curvature is good news: it turns our discrete search into something that admits pseudo‑gradients. Greedy steps, local improvements, neighborhood exchanges, and short look‑ahead policies—when guided by proper prices and fed with probabilistic inputs—climb quickly toward high ground. These problems are not cryptographic; flipping one bit does not turn feasible into impossible and collapse value to zero.138. Algorithms that rely exclusively on splitting abstract spaces and tightening linear relaxations tend to grind on the wrong geometry; the constraints of business are rarely tight enough, early enough, for half‑space pruning to pay.
Second, “the optimum” is conditional. Change the frame and the summit moves. Add a liquidation channel; allow price nudges tied to sailing dates; widen what may co‑load; legalize late binding between near‑substitutes: yesterday’s paragon becomes today’s local hill. This is not a failure of optimization; it is its purpose. In Kuhn’s terms, routine progress happens within a paradigm; step changes happen by changing the paradigm.139 A living decision engine must therefore be built for experimental optimization: dual‑run challengers against incumbents, promote only when measured deltas justify it, revert cheaply when they do not. The obsession with certificates of global optimality in toy worlds dies here; the currency is uplift measured on the firm’s books.
The solver within
A solver is the software component that, given a frame, forecasts, and prices, searches the admissible space and proposes commitments. It is the heart of the engine but not the engine itself. Around it sit the indispensable organs: semantic adapters to read authoritative records; probabilistic modules to produce distributions; valuation ledgers to expose prices and owners; halting heuristics to suspend emission when confidence drops; and the cut‑over machinery that promotes proposals to facts and re‑ingests them on the next cycle.
Because the engine runs under uncertainty and repeats, several engineering stances prove decisive.
First, stochastic control beats deterministic planning. The solver should rank admissible moves by their expected, risk-adjusted rate of return under the supplied probability distributions. Sequential dependence—today’s move shaping tomorrow’s options—can be handled by optimizing policies rather than single acts, typically within a receding‑horizon loop in which each cycle recomputes with fresher information. Functional forecasts—simulators that map “what we would do if…” to outcome distributions—are the natural partner for such policy search. The mathematics is straightforward; the novelty is to make it small, legible, and auditable enough to run every day.
Second, penalties tame unenforceable limits. In the stochastic world, “never exceed X” is either impossible or ruinous. Pricing violations brings limits under one roof and aligns them with the money objective. Chance-constraint language may be useful for discourse; prices do the work. We return to unenforceable limits—and to the relation between chance constraints and priced penalties—in what follows.
Third, design for repeat resolution. Most supply-chain choices are not one-off industrial design problems (e.g., siting warehouses or fixing lanes) but small, stochastic allocations made again tomorrow under slightly different facts. A solver that prospers here must ingest distributions, price options in coins, and be cheap to halt and rerun. The output is not a monolithic plan with a certificate; it is a steady stream of better wagers under a common ledger. This stance runs against the dominant OR posture—one-shot global models with static constraints—which is why only a handful of firms use solvers day-to-day in operations.
Fourth, instrument for trust. Every emitted decision must come with a dossier: the admissible options considered, the prices and distributions used, the marginal deltas that justified the choice, and the tests passed on the way out. Dual‑run should be the default: the incumbent recipe holds the pen while a challenger writes shadow decisions; measured differences accumulate in coins. When something smells wrong—tails miscalibrated, a constraint starting to bind—halt. Flooding users with “exceptions” deputizes them as unpaid coprocessors and teaches them to distrust the system. A clean stop invites the only interventions that help: edits to the frame, the prices, or the semantics.
Finally, a sober word about the toolbox sold under the OR label. After seven decades of research and two more of ever‑faster hardware, generic linear‑ and mixed‑integer solvers remain marginal in day‑to‑day supply‑chain operations. They show up for one‑off design work—network siting, long‑horizon capacity studies—and for a few tightly bounded niches. They almost never govern the recurrent, uncertainty‑laden choices that move coins every day. This is not managerial stubbornness; it is a mismatch between instrument and phenomenon.
Classical OR technology presumes a world that supply chains do not inhabit. Uncertainty is compressed into point surrogates or thin‑tailed “safety” terms; objectives are written in artifact units—service levels, accuracy percentages, utilization—rather than in money; and a monolithic plan is produced as if execution were a one‑shot plan rather than a stochastic control loop that repeats at operational cadence. The modeling creed that separates “pure statement” from “impure method” then forbids the very hints that practice requires: priced penalties for unenforceable limits, explicit shadow prices on shared resources, warm starts from yesterday’s commitments, and greedy‑accumulation rules steered by windows of responsibility. Faced with fat tails, ratchets, and option values, the global MILP either times out, optimizes the wrong thing, or yields a brittle schedule that collapses at first contact with reality.
The empirical resolution is simple and unglamorous. Treat solvers as subroutines inside a larger, economics‑aware, stochastic search rather than as sovereigns over the decision. Let probabilistic forecasts supply the futures; let the objective speak only in coins; move most “hard” limits into priced penalties; recompute marginal commitments at the cadence reality allows; halt when confidence drops; run challengers against incumbents in parallel; and embed domain structure wherever it buys return. Classical primitives—assignment, min‑cost flow, knapsack, routing—remain invaluable, but only when they are fed proper prices and glued into a loop that respects uncertainty, repetition, and ratchets.
Seen this way, the verdict on OR in operations is not that the mathematics is wrong, but that the posture is misaligned. Tools built for stationary toys and one‑shot plans cannot manage the combinatorics of options under fat‑tailed uncertainty where profit lives in the tails. Compute should therefore be spent where it purchases coins: on better frames, better prices, and better probabilistic views, not on certificates of optimality for the wrong question. The next section puts a price on compute itself.
The economics of compute
Optimization has a cost. Minutes of CPU, hours of wall‑clock, days of engineering—each must be justified the same way any other resource is: by its expected, risk‑adjusted return. The relevant horizon is operational. Rescheduling pick robots invites millisecond cadences; overseas buying does not. Vendors peddle “real‑time” as if time itself were a virtue; it is not. The right cadence is the one at which an extra unit of compute no longer repays itself in coins.
Two practical consequences follow. First, accept that some decision cycles are “slow by design”. Large, stochastic, high‑dimensional searches cannot be driven to hundred‑millisecond latencies without either gutting the frame or trivializing the objective. Amdahl’s law140 bites as soon as the engine must touch broad contexts to avoid local stupidity. Second, engineer graceful degradation. Run a quick pass that captures 80 % of the value with cheap heuristics and good prices; spend the remaining budget where shadow prices scream for attention; stop when expected uplift per minute falls below the minute’s cost. Because the engine repeats, tomorrow’s run will harvest today’s leftovers with fresher information and a warmer start.
The same calculus governs model size. Bigger instances are not “harder” in any metaphysical sense; they are only more expensive with a fixed technique. If a method blows past a sensible budget, change the method, coarsen the grain where economics permits, or shrink the horizon. Measured uplift, not aesthetic purity, decides.
In sum, optimization is the language through which a decision engine expresses preferences under uncertainty. When reframed around marginals, prices, and admissible moves; instrumented for repetition, halting, and audit; and engineered with sympathy for both hardware and economics, it becomes what the automation program needs: not an oracle that dictates a plan, but a tireless clerk of the firm’s intentions, forever proposing the next‑best wager and writing it—legibly—into the ledger.
Deciding under uncertainty
Decisions are priced wagers on futures we do not control. Earlier in this chapter, we adopted the stance that makes such wagers computable: the objective is deterministic and money‑denominated, while uncertainty enters through probabilistic forecasts—distributions for the facts that matter (demand, lead times, reliabilities, returns, sometimes prices). We do not bury randomness inside the objective; we feed the engine distributions and ask it to rank admissible commitments by their expected, risk-adjusted rate of return.
This posture departs from the teleological vision (Chapter The Future), which treats uncertainty as a defect to be engineered away by better point forecasts. Even small errors in such point estimates can have discontinuous consequences: a turbine grounded for one missing vane is worth far less than “99.98 % of a turbine,” and a rival that is 1 % cheaper can erase 100 % of your sales. In supply chains, profit and loss live in the tails. Methods that optimize against a single guessed trajectory therefore manufacture confidence in place of robustness.
Nothing here hinges on vocabulary. Calling the problem “stochastic optimization” simply acknowledges that futures are represented by distributions and that valuation is taken in expectation (and, when warranted, with explicit risk adjustments). What matters in practice is where uncertainty is carried and how it is priced. As argued above, the right place is in the forecasts that enter the objective, not in the objective itself; and the right prices are the firm’s own coin‑denominated penalties and rewards—stockout pain, obsolescence, capital charge per unit time, congestion, late fees, the option value of waiting.
One might be tempted to conclude that once expectations are taken, the “deterministic” and the “stochastic” cases collapse to the same mathematics. They do not, in practice. Deterministic recipes encourage brittle plans, ignore fat tails, and hide trade‑offs behind artifact KPIs; an engine designed for distributions must, instead, price tails, expose valuations, halt when confidence drops, and repeat at the cadence reality allows. This difference of posture—not a change of symbols—drives results.
With this foundation in place, the remainder of the section turns to operational consequences of deciding under uncertainty. First, we examine how point‑based methods systematically produce fragile decisions—choices that look lean on paper and then hemorrhage coins when the world wanders off the median. We then treat unenforceable constraints, which must be handled by prices or explicitly bounded risks rather than by edicts; we put a price on learning via the exploration–exploitation trade‑off; and we justify funding options that are unprofitable today but hedge tomorrow’s regimes. Each topic applies the same grammar: admissible moves, probabilistic views, and a single money ledger.
Fragile decisions
In the present book, a decision is fragile when a small deviation from the assumed “typical case” produces a disproportionate loss in its risk‑adjusted rate of return. Operationally, fragility appears as both a higher variance of realized returns and—once fat tails and ratchets are correctly priced—a lower expectation than competing moves that keep options open. Fragility is not about cataclysms; it is about the humdrum surprises that visit every flow: a supplier that ships short, a truck that misses a slot, a store that closes for a day, a post that goes viral and empties a shelf. Each event is surprising locally; taken in aggregate, they are not.
Fragility is largely manufactured by policy. The teleological posture—one ordained future, a plan to enact it, and performance indicators that police compliance—breeds decisions that look wonderfully lean on paper and hemorrhage coins when the world takes even a modest detour. Point forecasts compress futures to a single trajectory; “hard” limits are treated as edicts; utilization is pushed toward saturation; buffers are justified by formulas whose hidden prices no one owns. Such procedures do not merely face variability; they amplify its costs by removing the very options that variability makes valuable. It is therefore more precise to speak of fragile policies—rules that systematically produce fragile decisions.
Two structural features make these policies brittle. First, most supply‑chain payoffs are lopsided: the downside lives in the tails while the upside is bounded. One missing component can ground an aircraft worth millions; being 1 % dearer than a rival can wipe out the sale entirely. Thin‑tailed surrogates and median‑case planning hide these cliffs and thus underprice the option to avoid them. Second, operational ratchets erase alternatives. A premature store allocation that looked harmless under a baseline forecast prevents serving a better request tomorrow; a “full‑truck” routine that ships on the calendar locks tomorrow’s consolidation out. Decisions that maximize paper “efficiency” by binding early are, in practice, decisions that sell options cheaply and buy trouble dearly.
The optics are treacherous. Fragile decisions look efficient: trucks depart visibly full, distribution centers look empty, headcount looks thin, and inventory sits very close to the edge. Non‑fragile decisions look, at a glance, a bit “wasteful”: a sliver of capacity is kept available, a pool of stock is held back at the distribution center, a second supplier is retained, a few repairables are prepositioned. Yet once tails and ratchets are priced in money, the apparently conservative stance delivers a higher expected, risk‑adjusted aperiodic rate of return. The premium that buys slack is an option premium; it repays itself whenever the world misses the median—which it reliably does.
A mundane illustration clarifies the point. Consider line maintenance for a jetliner on a tight turn. The work scope is uncertain until the first panels open; one missing vane can strand the aircraft and cascade missed connections. A deterministic routine that provisions a kit sized to the “average job” will look lean ex ante and fail expensively whenever demand spikes on a few parts. A rugged routine begins with distributions for the uncertain work scope and for repair–return delays, attaches a coin‑denominated penalty to schedule slip, and prices each extra part as an option against that penalty. The resulting kit is not “minimal”; it is profitable. Unused parts go back to stock; their carrying charge is small next to the loss avoided when tails bite. The same logic governs any high‑penalty, low‑probability shortage: fat tails dominate and thin‑tailed arithmetic manufactures fragility.
A second illustration lies inbound. Suppose a warehouse faces random supplier drift and a finite number of receiving doors. A weekly, plan‑compliant ordering calendar will occasionally stack sailings and overload the dock; the remedy becomes firefighting—deferrals, driver detention and demurrage141, overtime crews, and off‑contract priority premiums for scarce dock slots.
A rugged stance treats arrivals and dock times as distributions, prices overload through explicit congestion penalties (the same economics that surface as detention, demurrage, overtime, or priority‑slot premiums), and lets the decision engine defer or advance orders until the expected, risk‑adjusted return of the next batch clears the firm’s shadow cost of capital plus the option value of waiting. The cadence becomes an output of economics, not an input from the calendar; the dock is not “always busy,” but coins stop leaking.
Downstream, premature allocation is another quiet source of fragility. Shipping scarce units early to weak stores freezes them where they rotate slowly and prevents serving the winners that emerge a few days later. A rugged routine attaches a visible shadow price to distribution center stock under pressure—coins per unit per day that represent the option to serve a better request tomorrow—and releases a unit to a store only when that store’s marginal return overtakes the distribution center’s hold price. Throughput usually increases where it matters most, even though more units sit at the distribution center for a little longer. The optics of “busyness” improve under fragile policies; the economics improve under robust ones.
Fragility also imposes a labor cost that escapes dashboards. Fragile policies collide with reality, generating long exception queues and conscripting planners as human coprocessors. The organization normalizes chronic expediting, manual resequencing, and after-the-fact overrides. This firefighting absorbs the attention needed to build the capital good at hand: a small, legible decision engine that encodes frames, prices, and probabilistic views, runs unattended, and compounds improvements. Fragile policies not only lose coins on bad days; they also keep the firm from investing the time that would make bad days rarer and cheaper.
Engineering non-fragile decisions does not require heroics; it requires a change in posture. The rugged view treats uncertainty as the raw material of profit. Practically, this means (i) admitting distributions—over demand, lead times, reliabilities, returns, and sometimes prices—as first-class inputs; (ii) expressing all frictions in coins—stockout pain, congestion, aging, late fees, the option value of waiting—and owning those prices company-wide; and (iii) letting a stochastic optimizer rank marginal commitments by their expected, risk-adjusted aperiodic rate of return. Small, priced slack follows naturally: a bit of pooled inventory at the distribution center, a few retained alternatives in transport, measured headroom at critical resources, modest over-provisioning of kits where tail penalties are sharp. These are not talismans against bad luck; they are options purchased because the expected payout exceeds the premium.
A simple test exposes fragility in practice. If a recommendation collapses under small, realistic perturbations of inputs; relies on calendar rituals rather than priced cadences; expresses “excellence” in artifacts—accuracy percentages, service targets, utilization scores—rather than in coins; and demands constant human “exceptions” to remain tolerable, it is fragile. Conversely, if a routine can articulate its frame, show its distributions, name its prices, defend its cutoff for action versus waiting, and produce dossiers that survive contact with realized outcomes, it is likely to be robust—even if, to the untrained eye, it looks a little “wasteful”.
Once uncertainty is admitted, one last ingredient becomes unavoidable: many limits cannot be guaranteed under all draws of the dice. Dock slots, driver rosters, pick faces, and battery charge cycles will occasionally be stressed by unlucky coincidences even when decisions are sensible. The right treatment is not to deny randomness but to price it. The next section turns to these unenforceable constraints and shows how to handle them with chance bounds—or better, with explicit penalties stated in the same coin that governs every other trade-off.
Unenforceable constraints
In a stochastic world, most “constraints” are not laws of physics but preferences backed by penalties. Treating them as edicts that must hold under every draw of the dice produces brittle plans and expensive firefighting. The frame introduced earlier already suggests the right split. A few limits are truly hard—legal prohibitions, physical impossibilities, safety barriers—and belong in feasibility tests that reject candidate moves outright. Everything else is soft: it can be violated for a price or delay and therefore belongs in the objective as an explicit, coin-denominated nuisance.
Let’s consider, for example, a company passing replenishment orders for a warehouse. The warehouse has a maximum inbound capacity. If too many trucks arrive on the same day, the staff must turn some away to return later. This is costly and disruptive, tying up trucks planned for later deliveries. Thus, replenishment orders are intentionally spread out to avoid excessive inbound shipments on any given day. However, suppliers may face difficulties. Unforeseen delays may, infrequently, cause deliveries—initially planned for distinct dates—to collide on the same day. Because supplier delays are random—and often fat-tailed—such clashes can never be ruled out. This would require impractically long delays between successive orders. Spacing orders mitigates the risk of exceeding inbound capacity, but it cannot eliminate it.
More broadly, most supply-chain constraints carry a real risk of violation even when handled carefully. By definition, the company has no control over its customers or its suppliers. Nor does it have complete control over its own employees, who may fail to show up for various reasons. Yet, while perfection is not of this world, the company can steer its decision-making to respect those constraints. Consider two approaches.
Chance constraints offer a less stringent criterion. Instead of insisting a constraint be met for all outcomes, a chance constraint requires it to hold with a minimum probability. One pragmatic approach is to allow a small, explicitly quantified risk of breaking a rule—for example, keeping the chance of dock overload below 5%. This keeps violations rare and controlled.
The main weakness of chance constraints, however, is that they are fundamentally noneconomic—or rather, that the economic trade-off is hidden inside an arbitrary percentage deemed acceptable—a dimensionless number. While one can tune this percentage to reflect the economics of the situation, doing so creates a new problem to solve.
Penalized constraints are a simpler alternative: move the constraints into the objective through penalties. They treat every breach as a costly nuisance and let the optimizer weigh that penalty against the decision’s expected benefit. Each violation incurs a large cost. The optimization then minimizes expected total cost, naturally encouraging feasible or near-feasible solutions. A sufficiently high penalty deters violations and steers the optimizer toward nearly feasible solutions.
Penalized constraints suit supply-chain optimization because they express operational limits in the same monetary terms that govern the model. Almost any bottleneck can, in principle, be relaxed for a price, and the penalty term captures that price explicitly. The optimizer therefore searches for plans that avoid ruinous outlays without being boxed in by rigid rules. Crucially, because each penalty is stated in currency, dozens of heterogeneous limits sum seamlessly into a single objective, allowing the model to scale gracefully in high-complexity environments.
Exploration vs. exploitation
Decisions repeat. Each cycle updates our beliefs, and those beliefs steer the next cycle. Historical data are therefore not neutral; they are the by‑product of yesterday’s choices. If a retailer has never priced a SKU above 9.99, the ledger contains no evidence about demand at 10.99. A solver that only “exploits”—always choosing today what looks best under the current model—quietly freezes the model’s blind spots. Profits then plateau not because the world is exhausted but because the firm no longer learns where returns might be higher.
The rugged stance (Chapter The Future) treats this endogeneity as a design constraint. Occasionally, we must pick a move that is slightly inferior on today’s ledger because the observation it generates will improve tomorrow’s ledger by more than the foregone coins. In economic terms, the firm buys a small option on knowledge. The practical question is not philosophical—whether to “try new things”—but quantitative: when does the expected value of the information exceed the opportunity cost of the experiment?
Two kinds of uncertainty matter. Aleatory uncertainty is intrinsic randomness; no amount of effort will erase it. Epistemic uncertainty is ignorance; it can be reduced by gathering the right observations (cf. Chapter The Future). Exploration is the deliberate reduction of epistemic uncertainty when its expected dividend—better decisions over the relevant horizon—outweighs the immediate shortfall of the probe.
A small, concrete illustration makes the endogeneity point plain. Consider a sports‑drink line sold across one hundred stores. The firm has long held an informal doctrine that “ten-coin” is the sweet spot for impulse purchases; as a result, the price database is rich around 9.99 and almost empty above 10.49. The demand model inherits this bias and, when asked, advises once more to stay near 9.99. A rugged routine counteracts this inertia without gambling the brand. It designates a tiny test cell—say, eight stores with typical footfall—for two weeks, raises the price to 10.49 there, and re‑estimates the local price response. The cost is the estimated margin lost (or customers deterred) against the status quo price during the test period; the dividend is the uplift the new response unlocks when rolled to the other ninety‑two stores. If the expected dividend, stated in coins, exceeds the experiment’s cost, the price probe is rational—even if the eight stores underperform during the test.
Exploration is not confined to pricing. A buyer facing a new supplier with uncertain reliability can place a handful of deliberately small, early orders to learn the tail of the lead‑time distribution before committing volume. A network planner can try a modest cross-dock scheme in one region to measure congestion penalties rather than assuming them. A merchandiser can let a few low‑sell‑through items stock out on purpose (on a safe scale) to learn true substitution rates instead of extrapolating from censored data. In each case, the engine prices the probe like any other move: admissible options are stated, futures are probabilistic, and all effects—shortage pain, aging, congestion, capital charge per unit‑time—are expressed on the same coin ledger.
This trade‑off is known in the literature as “exploration vs. exploitation”.142 The classical bandit imagery is useful, provided we keep the economics in view. The goal is not to chase abstract “regret bounds” but to decide, at operational cadence, whether a marginal probe clears the same bar that governs every commitment: expected, risk-adjusted aperiodic rate of return. Three design habits keep the process legible and auditable:
First, keep probes small and bounded. The window of responsibility limits attribution; scope (number of stores, lanes, SKUs) limits downside. Smallness is not timidity; it is how we price and compare learning options cleanly against their dividends.
Second, make the learning dividend explicit: the expected uplift—over the window during which the revised rule will be in force—net of aging, markdowns, congestion, and capital charge. Hidden “benefits” count as zero; only coins on the ledger count.
Third, embed exploration within the same unattended engine. A probe is not a special project; it is another admissible move with a dossier: the prior, the proposed perturbation, the window, the prices, and the rule for promotion or rollback once outcomes arrive. If the engine’s confidence drops, halting heuristics suspend the probe exactly as they suspend ordinary emissions.
This stance also clarifies why the teleological view cannot accommodate exploration. Under forecast‑plan‑comply, decisions are only the last line of a deterministic program; forecasts live apart from actions; and “accuracy” is policed as a percentage detached from coins. In such a world, there is no budget for buying information because the procedure presumes it already knows what matters. In the rugged world, by contrast, quick, priced probes are a source of edge: they narrow epistemic uncertainty where it pays and let the firm move first where a small insight compounds.
We now turn from buying information to buying insurance. Exploration funds observations that change the model; nurturing options fund capabilities that are unprofitable today but would pay—handsomely—under specific regimes. The grammar remains the same: admissible moves, probabilistic futures, one coin ledger.
Nurturing options
Resilience means funding options that are unprofitable today yet could preserve future profits if conditions shift. Indeed, whether a given range of options—offered by a supplier, a carrier, or an assembly line—is valuable to the business depends on prevailing market conditions. If those conditions change, certain options may sharply appreciate or depreciate. The company can hedge such risks by nurturing the right options.
For example, a company based in Westonia may presently source a part from a supplier in Ruritania. Local alternatives in Westonia are, unfortunately, not competitive with the Ruritanian supplier. Moreover, their parts do not fully meet the company’s specifications. The company is fully satisfied with its Ruritanian supplier. Yet Westonia’s government may impose a massive tariff on imports from Ruritania. Indeed, several rising political figures have been outspoken in favor of such tariffs. If enacted, such tariffs would severely reduce profits, and the firm could not switch to local suppliers because their parts fail on specifications. Thus, to insure against this risk, the company selects a local supplier and helps upgrade its production so the parts meet specifications. Most orders still go to the Ruritanian supplier, yet the firm nurtures the local alternative, which may prove a profit-saver if the tariff is imposed.
Nurturing options take many forms. The company might fund unused capacity—manpower, storage, production, or transport. The company might pay high service fees in exchange for favorable termination clauses (to surrender a lease of a vehicle, a building, or a piece of industrial equipment). The company might lower assembly-line efficiency to increase reconfigurability, broadening the range of designs the line can output.
The point of nurturing options is to prepare the company for conditions that are unlikely yet far from impossible. Unlike the exploration–exploitation trade-off discussed previously, the investment’s goal is not discovery. From the outset, the option is understood to be nonviable so long as present market conditions persist. The investment is a hedge.
Assessing viability requires quantifying three elements: (1) the probability of the trigger, (2) the savings the option would yield under those conditions, and (3) the cost of nurturing it. If probability-weighted savings exceed nurturing costs, hedging is profitable.
Again, nurturing options make sense only under the rugged view. Under the teleological view, the future is treated as a known quantity. As a result, options not immediately economically viable are dismissed. Indeed, without a supporting rationale, management struggles to justify “unnecessary” expenses—especially those tied to highly pessimistic conditions that would invalidate the current course favored by the management team.
Sequential decisions
Many supply-chain choices are sequential: the payoff of a move taken today depends on moves the firm will choose tomorrow. Present commitments erase alternatives and reshape next cycle’s option set—a ratchet effect noted in “Asymmetry of time” (Chapter The Future). The apparent paradox—“how can I optimize now when tomorrow’s decision is undecided?”—is routine: overseas buying that must fill a container, store dispatches that preempt later allocations, production batches that tie up capacity until the next changeover.
In this chapter, we treat such situations not as isolated one-off optimizations but as rules for acting over time. The clean formal object is a policy.
Reifying a policy does two things. First, it brings the frame into view: admissible moves and feasibility tests; the money-denominated prices (penalties and rewards) the firm is willing to defend; and the probabilistic view of futures that informs each choice. Second, it enables closed‑loop evaluation via a functional forecast—a simulator that answers, “if we follow this policy, what distribution of outcomes should we expect?” This pair—policy plus functional forecast—lets us compare candidate rule‑sets by the same yardstick used for one‑shot decisions: expected, risk‑adjusted rate of return.
That said, turning every sequential problem into a full policy search rarely justifies the opacity and compute it entails. The space of possible rules is vast; scoring them only at the level of whole trajectories obscures marginal attributions; and naïve look‑ahead becomes too expensive for day‑to‑day operations. In practice we keep the marginalist stance—rank small, auditable commitments by coins per unit time—while using a few disciplined instruments that capture the sequential coupling without reifying a full policy.
A running example fixes ideas. Suppose a buyer must raise an overseas purchase order that is economical only when rounded to a full container. The mix can combine many SKUs, but every unit loaded today must cover demand until the next container from the same supplier arrives. Yet that arrival date depends on a decision the firm has not yet made. A literal, single‑shot optimization cannot value today’s order without assuming tomorrow’s. A policy resolves the dependency (“order when the shadow price of the next container exceeds $$S$$; allocate units by marginal return subject to weight/volume constraints”), and a functional forecast projects the distributions of lead times, demands, and dock capacities under that rule.
The remainder of this section shows how to handle most of this coupling with lighter means: a window of responsibility that bounds the horizon over which today’s order is held to account (next subsection); the economics of waiting, which prices the option to defer and sets a cutoff rate of return for action; permutation invariance, which justifies greedy accumulation of batched allocations—with caveats where order truly matters (vehicle routing, job-shop scheduling); and the postponement principle, which favors late binding to keep future options open. Only when latency or scale demands it do we perform an explicit, parameterized policy search.
Window of responsibility
Sequential decisions create an attribution problem: today’s move reshapes tomorrow’s options, and tomorrow’s move, in turn, will repair or amplify today’s consequences. An engine that must run every day needs a rule that assigns consequences to each commitment without simulating the entire future. The practical device is a window of responsibility: a bounded horizon over which the present decision is held to account, while later decisions inherit the rest.
The window is not a planning bucket; it is an accounting principle. It answers two questions: from when do we start attributing outcomes to today’s commitment, and until when does this responsibility last before it is handed over to the next decision of the same class?
For an overseas purchase order that must fill a container, the natural start is not the calendar day the order is entered, but the day the goods become usable—after production and transport—because only then can the decision begin to earn or lose coins. The end is the first later moment after which a subsequent order can reasonably take over responsibility. In practice, a robust convention is to end the window at the arrival of the next–next container from the same source. The immediate “next” container is still partly correcting or amplifying today’s choice; after the “next–next”, responsibility has genuinely migrated. By contrast, for daily store replenishment, the window typically runs from tonight’s delivery to tomorrow night’s; the cadence itself supplies the boundary.
Within the window, the decision is scored in money terms consistent with the firm’s objective. Sales (or consumption) realized during the window are credited; shortages that occur within the window are charged according to an explicit, coin-denominated penalty that reflects lost margin, goodwill, and any cascading effects on adjacent items. Inventory left at the end of the window must not be swept under the rug. Two treatments are practical and, depending on the case, equivalent in spirit. The first is a terminal valuation: treat the residual stock as if it were immediately liquidated at its expected salvage value, less the expected carrying and handling costs to reach that liquidation. The second—often simpler and more conservative for long tails—is to charge the entire expected stream of future holding and obsolescence costs to the decision that created the stock. This “ratchet” treatment recognizes that later choices cannot undo an over-large purchase without paying for markdowns, storage, and write-offs; the load rightly stays with the originator.
Defining the window’s bounds is a probabilistic exercise. Two distributions usually suffice. First, a lead‑time distribution places mass on the dates when the commitment begins to matter. Second, a distribution for the next opportunity to re‑decide marks the hand‑over date. This second forecast is “impure” in the sense that it projects our own future behavior; yet at the batching scales that matter—container sailings, factory changeovers, weekly dispatch slots—the cadence is largely governed by externalities (cutoffs, vessel schedules, MOQs) and proves stable enough to forecast with useful accuracy. Where cycles are rigid—nightly store dispatches, end‑of‑shift production releases—the calendar itself supplies the end.
Truncating the horizon obliges us to price what lies beyond it. The stockout penalty already mentioned is one example: it stands in for repeat purchase erosion and substitution dynamics that we choose not to simulate across seasons. Similar shadow valuations are sometimes needed elsewhere. A fashion brand that discounts aggressively at season’s end may find that heavy rebates teach customers to wait; the long‑run demand damage is real, yet difficult to model across future collections that are not even designed. A small, per‑unit discount malus—expressed in coins and applied inside the window—keeps the engine honest without turning the simulator into a multi‑year saga. As always, prices must be explicit, owned, versioned, and exposed; hidden penalties do not become neutral—they become capricious.
A brief illustration fixes ideas. Suppose an importer orders a full container today. The evaluation window starts when the container clears and goods are available; it ends at the arrival of the next–next container from the same lane. All sales made in between are credited to today’s order; all stockouts in the same period are charged according to the shortage penalty. Residual units at the window’s end receive a terminal valuation—or, in the ratchet approach, the expected future holding and markdown costs are fully charged to today’s order. The next container is evaluated under its own window, with the same rules; responsibility does not double‑count. This convention makes the daily engine tractable and auditable while preserving economic realism where it matters: in the tails and in the irreversible costs that later choices cannot erase.
The window of responsibility harmonizes with the marginalist stance developed earlier. It lets the solver reason one step ahead with probabilistic inputs, attribute consequences cleanly, and move on. Its validity is empirical, not theoretical: choose the boundaries, run dual‑mode against incumbents, measure the deltas in coins, and adjust. In practice, this device succeeds across a wide range of sequential settings—overseas buying, warehouse inbound smoothing, store dispatches—because it keeps the focus where economics lives: on priced options and fat‑tailed futures, not on speculative multi‑year narratives.
The economics of waiting
Every sequential problem contains a silent option: do nothing yet. Waiting preserves options and lets information accumulate; acting now locks in a rate of return and erases alternatives. The engine should price delay explicitly and adopt a cut-off rule. Act only when the expected, risk-adjusted aperiodic rate of return of the best admissible move exceeds the shadow cost of capital plus the option value of waiting. Below that bar, defer.
Operationally, this becomes a simple, repeatable discipline. Recompute, daily or at a reality-driven cadence, the best next move within the current window of responsibility; if its rate of return clears the threshold, take it and record the commitment; otherwise, wait. After a container has just been ordered, the very next day’s container is usually unprofitable: demand has not yet accumulated, consolidation opportunities are thin, and inbound congestion penalties loom. As days pass, volumes aggregate, co-loading improves, and the expected return rises; once it crosses the bar, the order is placed. The policy is not “weekly cycles by tradition”; it is “whenever economics justifies it”.
The same logic governs downstream dispatch. Consider a fashion DC holding a handful of units of a runaway bestseller. Allocating one unit to a weak store today may look harmless under a point forecast; tomorrow it becomes an expensive ratchet when a strong store asks for it and none is left. Instead of attempting to script the entire season, attach a shadow price to DC stock under pressure—a coin‑per‑unit‑per‑day opportunity charge that represents the option to serve a better request later. Release to a store only when that store’s marginal return—computed under the current probabilistic view of sell‑through and shortage penalty—overtakes the DC’s shadow price. Far from hurting throughput, such patience tends to increase it where it pays most; premature allocation often reduces realized throughput by freezing stock where it will rotate slowly.
Waiting is not dithering; it is a priced decision like any other. The cut-off embeds both the firm’s cost of capital and its tolerance for tail risk. It also carries the halting heuristics discussed earlier: when the engine’s confidence drops (miscalibrated tails, unusual constraint binding), emissions pause until the frame, the prices, or the semantics are repaired. In this way, the window of responsibility and the economics of waiting combine to turn many sequential problems into a one-step greedy accumulation that stays aligned with profit. The next subsection, on permutation invariance, shows why such greedy accumulation often works—and where order truly does matter.
Permutation invariance
Many supply‑chain allocations are enacted in batches: a container is filled with a mix of SKUs; a line produces a run of units; a truck leaves the DC loaded for several stores. In these settings, what matters economically is the final multiset of marginal commitments in the batch and the constraints it satisfies (weight, cube, slots, cut-offs)—not the notional order in which the atomic commitments were picked. When feasibility tests depend only on totals and penalties and rewards are priced symmetrically for the items in the batch, the objective function is—up to negligible bookkeeping effects—invariant to any permutation of the micro-allocations that compose the batch.
This observation is powerful when combined with the instruments introduced earlier. The window of responsibility bounds the horizon over which a batch is held to account; the economics of waiting supplies a cut‑off: act only when the best admissible move clears the shadow cost of capital plus the option value of delay. Given both, the batch can be assembled greedily: at each step, recompute shadow prices for the scarce resources and pick the next admissible unit with the highest expected, risk-adjusted aperiodic rate of return; stop when capacity binds or when the cut-off fails. Because the batch valuation is insensitive to the permutation of its constituents, this stepwise construction targets the set that matters (the content of the batch), not a ceremonial sequence that has no economic standing.
Two caveats clarify what is—and is not—claimed. First, while the final value of a batch is order‑insensitive, the marginal value of any given unit is path‑dependent: early picks face laxer constraints than late picks. This is precisely why the greedy accumulation must recompute shadow prices after each addition; the path dependence is handled by repricing, while the set that emerges remains the goal. Second, permutation invariance is not a proof of global optimality; it is the engineering rationale for a search that climbs quickly toward high ground in problems dominated by diminishing returns and priced penalties—exactly the geometry discussed earlier in On optimality. The engine repeats at an operational cadence; tomorrow’s run starts warmer and narrower, and the window mechanism attributes consequences without simulating the entire future.
Concrete cases make the point. When filling an overseas container, feasibility cares about totals (weight/volume) and cut‑off dates, while value is earned by the mix that will sell within the evaluation window. It is sensible to add units—possibly across many SKUs—in decreasing order of marginal return, pausing whenever the economics of waiting says “not yet”, and resuming as demand and co-loading opportunities accumulate. In DC‑to‑store dispatch, the DC stock carries a shadow price that reflects the option to serve a superior request tomorrow (cf. The economics of waiting). A store receives a unit only when its marginal return overtakes that DC shadow price; within a departure wave, units are then added greedily until the truck’s priced penalties and capacities advise stopping. The result is often a higher realized throughput where it matters most. Premature allocations freeze stock in low-yield locations, a bustle that depresses throughput once the season reveals genuine winners.
Permutation invariance also helps keep models small and auditable. Batch constraints—MOQs, full‑truck incentives, dock congestion—are better expressed as feasibility tests and priced penalties over the set that ships than as procedural rules that script a picking order. The former yield clear prices and clean halts; the latter breed artifacts and exception handling. Because the batch’s value does not depend on a ceremonial sequence, the solver is free to propose the next best marginal commitment, log its dossier, and repeat until the cut‑off bites.
Order-sensitive allocations
Not all sequential problems admit permutation invariance. Some classes are intrinsically order-sensitive: the route of a vehicle, the sequence of jobs in a shop with changeover times and scarce skills, the traversal of a picker through an automatic store. There, the distance traveled, the queuing and blocking, or the changeover losses depend directly on the ordering of tasks. Swapping two stops in a milk‑run can add an hour of driving; permuting two steps in an MRO work‑scope can lengthen turnaround time by a day. In such cases, the sequence is economically real, and the model must say so.
The methodology, however, remains consistent with this chapter. We keep the frame explicit (admissible moves and feasibility tests), keep the economics open (priced penalties for lateness, overtime, missed windows, and aging), and ingest probabilistic views of demand, travel times, repair durations, and reliabilities. The solver’s job is then to propose a sequence, not merely a set. Windows of responsibility still bound attribution—end of shift, end of day, return‑to‑depot—and priced penalties bring unenforceable limits into the objective. The principle remains: make order explicit only where it changes money; elsewhere, treat it as an artifact and let permutation invariance justify greedy accumulation.
Seen together, permutation invariance and the economics of waiting encourage late binding. When the valuation of a batch depends on its contents but not on the ceremony of their selection, it is prudent to defer commitments until the marginal returns clear the threshold, and then assemble the batch quickly by greedy accumulation. The next section, the postponement principle, formalizes this stance and shows how to design flows so that binding choices are taken as late as practicable, preserving options and improving returns under uncertainty.
Postponement principle
Permutation invariance and the economics of waiting point to a general rule for sequential work: bind late. The firm earns its edge by deferring irreversible choices until information arrives that materially changes the ranking of options, while charging a fair price for the delay. Under the teleological vision (Chapter The Future), deferral is a nuisance that complicates plan compliance. Under the rugged vision, deferral is a priced option: it preserves alternatives and is exercised only when its expected, risk‑adjusted return clears the relevant bar.
Postponement—also called late binding—means arranging the flow so that binding choices are made no sooner than the economics justify. Keep goods, capacities, and commitments in generic, reassignable states until the expected risk-adjusted aperiodic rate of return on binding exceeds (i) the shadow cost of capital, (ii) the option value of waiting, and (iii) the postponement premium—the incremental costs paid to keep options open.
The premium is tangible: carrying semifinished inventory that can be specialized later; keeping retainers with alternate suppliers; reserving dye-house or kitting capacity; paying extra handling to keep stock pooled at the distribution center; and investing in more sophisticated orchestration. Postponement is not “free flexibility”; it is a bet that the value of information and optionality will more than repay the premium and any aging risk within the window of responsibility. Properly applied, postponement tends to shorten the half-life of decisions (see Changing course below) by moving commitments closer to the moment they start to pay back and by reducing costly misallocations that would otherwise linger.
Late binding takes several recurring forms, all governed by the same economic test.
Form postponement. A fashion importer may commit early to greige fabric and trims (cheap, generic inputs) and defer the expensive specializations—dye, wash, label, size mix, geographic split—until early sell‑through reveals which articles and colors carry fat tails. The admissible options become explicit: finish now and ship broadly, or wait a few weeks and finish what the winners will actually absorb. The decision engine prices the salvage value of the laggards, the shortage penalty on the winners, the specialization costs, and the premium paid to keep the dye-house on call; it then binds only when the expected aperiodic rate of return clears the cut-off.
Allocation postponement. Shipping scarce units to stores too early freezes them where they rotate slowly. Holding them at the DC with a visible shadow price (coins per unit per day) preserves the option to serve a superior request tomorrow. A store receives a unit only when its marginal return—computed under the current probabilistic view of sell‑through and shortage penalty—overtakes the DC’s hold price (see The economics of waiting). Throughput typically increases where it matters most, because winners are fed first and laggards do not hoard.
Finish/kitting postponement. Coloring, kitting, packaging, or regionalizing variants early commits the firm to a demand split not yet known. Keeping WIP generic and finishing late lets orders reveal the mix. The premium is the extra handling and the capacity retainer at the finishing step; the benefit is fewer write‑offs on the wrong variants and a shorter decision half‑life as stock is specialized only when a taker exists.
Temporal postponement. On inbound lanes, a container ordered “because it’s Tuesday” fails the economic test more often than not. As days pass, volumes accumulate, co‑loading improves, and inbound congestion penalties evolve. Recompute daily; issue the order on the day the best admissible batch clears the threshold defined by the shadow cost of capital, the option value of waiting, and the postponement premium. The cadence becomes an outcome of economics, not of the calendar.
Price postponement. Prices shape demand; postponing markdowns or promotions until uncertainty has narrowed can dominate a scripted price path. Functional forecasts (Chapter The Future) make the pricing policy explicit and let the engine compare “markdown now” with “wait one more week” under the same money ledger. The premium here is the risk of being left with aged stock if the window closes; the benefit is extracting more from the winners without teaching customers to wait.
Operationally, postponement is the art of unbundling commitments into decisions that can be bound late. Break the monolith (“all articles, quantities, destinations, dates decided nine months ahead”) into staged micro-decisions that match the flow’s physics. Commit early to low-information, high-lead-time inputs; defer the high-information, low-lead-time specializations; keep stock pooled until assignment pays; and contract for flexible volumes and late specifications where the premium is tolerable. Managers often “simplify” sequential problems by rigidifying them—freezing assortments, banning mid‑season reallocations, locking price paths—not because physics requires it but because attended processes cannot cope. With unattended systems of intelligence, the bottleneck should sit in atoms and hours, not in the decision stack; late binding then becomes not only feasible but routine.
Like any heuristic, postponement has limits of jurisdiction. Where order is intrinsically consequential—vehicle routing, job-shop sequencing, picker traversal—binding early is the problem, and sequence must be optimized rather than deferred. Regulatory pre-clearance, scarce changeover skills, or perishable goods can also set hard fronts beyond which waiting no longer pays. The remedy is not to abandon postponement but to push the late-binding frontier upstream as far as economics and feasibility allow, and to price the remainder explicitly in the objective: penalties for lateness, aging, overtime, or missed windows.
In practice, the rule is simple and repeatable. At the cadence dictated by reality, recompute the best admissible binding move under the current window of responsibility. Bind only when its expected, risk-adjusted aperiodic rate of return clears the sum of the shadow cost of capital, the option value of waiting, and the postponement premium. Otherwise, wait. Where late binding cannot be engineered or where latency must be kept in the tens of milliseconds, precomputed, parameterized rules are warranted—the topic of the next subsection, policy search.
Policy search
The previous pages handled sequential decisions without reifying a “policy”. A window of responsibility bounded attribution; the economics of waiting priced inaction; permutation invariance justified greedy accumulation; late binding preserved options. In many flows this is sufficient and superior: it keeps the grain marginal, the prices explicit, the reasoning legible. Yet there are cases where we must compress computation into a rule that can be applied at millisecond cadence, or where the physics of the problem makes order intrinsic rather than ceremonial. In those cases we do not abandon the economic stance; we codify it into a policy and search its space.
A policy, as defined earlier, is an algorithm that maps states to admissible actions under stated prices and probabilistic forecasts. Turning that definition into practice serves a single purpose: to shift heavy computation upstream—in an offline, stochastic search—so that inference at decision time is fast and dependable. Where latency is generous (minutes to hours), recomputing the best marginal commitment with fresh forecasts is usually better than carrying a policy. Where latency is tight (tens of milliseconds), or where thousands of independent agents must act concurrently under local information—picker fleets in an automated warehouse, dynamic slotting inside a shuttle store, order-sensitive job sequences with scarce changeover skills—a well-designed policy becomes the right vehicle for the economics already developed in this chapter.
Beware the lure of “stationary proofs”. The literature is rich in analytical policies that are “optimal” only in toy worlds—stationary demand, thin tails, additive costs, no ratchets. Our world is rugged: distributions shift, tails dominate, and commitments erase options. Optimality in symbols should never outrank measured uplift under a faithful simulator and, later, in dual-run production.
When, then, is policy search warranted? Whenever order truly changes money and cannot be recomputed fast enough; whenever computation must be pushed off the critical path without sacrificing economics; whenever thousands of small agents must act coherently from local views; and whenever a simple, parameterized rule can carry the spirit of postponement, marginal valuation, and priced constraints with orders of magnitude less latency. When these conditions do not hold, the lighter instruments already introduced—windowed attribution, the economics of waiting, permutation invariance, and late binding—remain preferable: they move the bottleneck back to atoms, keep options open until binding pays, and avoid encasing yesterday’s beliefs in rules that will age.
Properly placed, policy search completes the picture rather than displacing it outright. It is how a firm converts deep, slow computation into fast, repeated action while preserving the same economic ground: decisions remain marginal wagers, priced in coins, justified under probabilistic futures, and revised as reality reveals itself. The next section turns from rules to inertia and quantifies how burdensome a commitment really is—its load and its half-life—so that, between two equally profitable moves, the firm prefers the one that lets it change course sooner.
Changing course
In a free market, the only constant is change. Being able to change course is a matter of survival for any company. Yet any supply chain decision comes with a ratchet effect: once resources are allocated, they are no longer immediately available for other uses. Some decisions appear final, such as a manufacturing process consuming raw materials, while others appear reversible, such as moving stock between warehouses. In any case, from an economic perspective, those decisions can be undone, though the required time and money vary.
When considering decisions with equal projected rates of return, a company should favor those with smaller underlying commitments, in time or in money. Avoiding unnecessarily large commitments is essential to keeping the company agile and, through agility, to withstanding the market shocks that will inevitably come.
Under the teleological view—forecast, plan, comply—there is no operative notion of agility at all. The future is treated as a schedule to be enforced, and the instruments gravitate toward plan-policing indicators: forecast accuracy, service-level percentages, utilization, inventory turns. Such gauges say little about the firm’s ability to change course; they reward adherence to a guessed trajectory and remain blind to the time required for commitments to unwind. Agility, however, is the capacity to shed or reassign commitments quickly when new information arrives, at tolerable cost.
To discuss agility in economic terms, we must replace plan-compliance surrogates with measures that price commitments over time. The classical trio—lead time, working capital, and inventory turnover—remains in view, but we recast them through two operational quantities that make duration explicit: the load a decision imposes (money $$\times$$ time) and the half-life of that load (the time required to release half the resources committed by that decision).
Lead time, in a broad sense, is the delay between a decision and the renewed availability of a resource. Working capital is the difference between current assets and current liabilities143. Inventory turnover measures the number of times inventory is sold or used during a year. Taken together, these indicators tell us how much capital is tied up but largely ignore for how long; the forthcoming notions of load and half-life address that blind spot.
All else equal, shorter lead time, lower working capital, and higher inventory turnover increase the capacity to change course. Indeed, these concepts characterize the company’s general inertia: its incapacity to change course immediately, as it must first liquidate the mass of assets presently flowing through while preserving their value.
However, these concepts are of limited use in understanding the company’s inertia. This is not to say these concepts are wrong, or even misguided; rather, the matter—the characterization of the company’s inertia—admits alternative concepts that are operationally superior.
The load of a decision
The (financing) load of a flow decision quantifies the underlying resource commitment as the product of a resource’s value and the lifespan of the goods it creates, with units of money $$\times$$ time.
Consider a simple illustration. If the company buys a batch worth 10 coins and resells it 5 days later, the load is $$L = 10 \times 5 = 50$$. Alternatively, if the company were to sell half the batch 5 days later and the other half 20 days later, the load is $$L = \frac{10}{2} \times 5 + \frac{10}{2} \times 20 = 125$$.
The second situation yields a load more than twice as large as the first because the company holds inventory for more than twice as long.
Now suppose payment for the batch is deferred by 12 days. The load of the first situation is effectively zero, as no financing is required: the supplier is paid after the goods are sold. The load of the second situation is $$L = \frac{10}{2} \times (20 - 12) = 40$$.
Deferred payment can substantially reduce the financing pressure that sustains the flow.
Load measures the working capital tied up over the entire life of a decision, not just at a reporting date. Whereas ordinary working-capital figures give a snapshot, load attaches a duration to every coin committed. For a firm under financing pressure, this time-weighted metric—read alongside expected return—is the sharper guide for ranking options and steering scarce capital to its best uses.
Many firms require managerial approval once a purchase order passes a threshold. The rule checks size and potential liquidity impact, yet it overlooks the goods’ lifecycle. A small order can immobilize capital for months, while a large backorder may clear overnight. Load captures this hidden, time-based burden directly.
Firms often plot inventory value against turnover to spot capital traps, but load already fuses both into a single metric.
The half-life of a decision
The half-life of a decision is the elapsed time—typically measured in days—until its load halves through sales, consumption, or service.
Consider an overseas order where delivery takes 10 weeks and a full container lasts 20 weeks before half the stock is sold. Thus the half-life is $$7,(10+20)=210,\text{days}$$.
Borrowed from physics, half-life measures the time for a decision’s load to halve; it supplies a time dimension that lead time alone obscures. Indeed, lead time understates the consequential duration of a decision whenever batching is involved. A similar critique applies to cycle time and inventory turnover. Naturally, the “half” threshold is arbitrary; any quantile could be used instead. Revisiting the example, the 75%-life of the purchase would be $$280$$ days, assuming constant demand.
Half-life gauges how quickly a decision can be unwound; long half-lives trap the company on its current course and erode agility.
Moreover, half-life resists manipulation better than lead time: gaming an indicator—making the number look good without improving reality—becomes easier when the metric ignores how long capital remains locked. This often stems from misaligned incentives that reward the indicator itself rather than the outcomes it is meant to represent. Indeed, companies seeking agility often tackle the problem by compressing lead times. However, a frequent pitfall is reducing lead times at the expense of half-lives.
Engineering
Modern supply chain practice operates with at least one degree of indirection: people engineer decision-making processes, and those processes emit myriad routine decisions that steer the flow. The emitting artifacts are numerical recipes: programs that ingest authoritative records, carry explicit economics, and issue commitments that can be enacted and audited144. In preceding chapters—Information, Intelligence, and Decisions—we argued that such engines must run unattended if the firm is to turn judgment into a productive asset rather than an endless clerical burden. This chapter turns to the engineering principles that let these engines exist and compound.
The scope is narrow by design. We will not rehearse generic software topics—UI145 frameworks, authentication, CI/CD plumbing146, control planes, message buses, or the wiring to and from systems of records. There is ample literature on such auxiliary sciences. Our concern is the junction where economic reasoning meets code: how one frames admissible options, carries uncertainty without flattening, prices trade-offs in money, and preserves auditability so a decision’s lineage remains legible months later. The tone remains deliberately non-technical—no algebra, no code listings—so we can be clear about the foundational choices that make a system of intelligence emit unattended decisions safely, repeatedly, and profitably.
Two clarifications anchor what follows. First, “numerical recipe” denotes the whole apparatus that matters in practice: not only predictors and optimizers, but also the instrumentation that reveals why a recommendation is sane, the halting conditions that stop emissions when confidence falls, and the dossiers that let practitioners understand, in money terms, what the engine believed at the time. Second, while the application landscape of every firm is idiosyncratic, the engineering stance advocated here is general: prefer explicit frames over tacit conventions, probabilistic views over point surrogates, priced penalties over hard edicts that cannot be enforced, and reproducibility over heroics.
LLMs warrant a brief comment. Their role is increasing and will become critical as they compress the cost of drafting, refactoring, testing, and documenting numerical recipes. Yet they are not a substitute for technical proficiency. LLMs do not supply economic framing, resolve data semantics, or enforce determinism and auditability; they merely speed the hands that perform the work. Treated as able scribes—instrumented, reviewed, and constrained—they shorten the iteration loop that experimental work requires. Treated as oracles, they invite subtle, costly errors.
Finally, a call to action: competitive advantage no longer rests on a thicker stack of dashboards—assuming it ever did—but on owning a small, engineering-minded capability that writes the firm’s economics into software and lets it run unattended. The next section explains why this capability cannot be purchased as a packaged “planning” layer and why, for systems of intelligence, programmability is not a taste but a necessity for safety and profit.
Why program
Before proceeding, a scope note. In Chapter Information, we distinguished systems of records, systems of reports, and systems of intelligence. What follows concerns the last class. Ledgers and workflow engines—ERPs, WMSs, CRMs—are packaged successfully because they record discrete events and enforce rule-based routes; entities are largely decoupled by construction. Configuration can express the modest couplings that workflows introduce, and COTS (Commercial off-the-shelf) suffices. What resists packaging is the engine that composes those records into unattended, economically priced decisions.
The emergence of enterprise software in the late 1970s extended the postwar hope—nurtured by operations research—that the quantitative problems of supply chain would admit a closed, permanent formulation. Vendors and academics aimed to crystallize those resolutions into reusable frameworks so that decision-making could be “configured” once and for all by non-technical users. This ambition, sensible for ledgers, was transplanted wholesale into systems of intelligence.
Half a century on, this promise has not materialized for systems of intelligence. COTS “planning” stacks take months or years to install, mobilize armies of specialists, and still degenerate into spreadsheet workarounds when they meet the grain of a living flow. The failure is structural—a mismatch between generic templates and business-specific couplings.
Closed analytical resolutions are brittle because they presuppose that what must be decided can be fully enumerated in advance. Reality refuses to comply. The only durable advantage that spreadsheets enjoy—despite their many defects—is precisely their programmability. They let practitioners encode idiosyncrasies the moment those matter economically. Programmability is not a convenience; it is the essential property of a system of intelligence.
Innocent‑looking quantities expose the brittleness. Consider bulk cable. A ledger that says “10,000 meters on hand” hides what actually governs feasibility and value: composition and cut. Ten coils of one kilometer and a thousand coils of ten meters both sum to 10,000, yet they are opposite economic assets once customer length profiles, waste on cutting, and residuals are priced. The unit—meter—obscures the state that decisions must reason about.
Nothing about cable is exotic. Similar, perfectly ordinary quirks pervade every vertical. In aviation, rotables and repairables exhibit asymmetric compatibilities—parts that substitute one‑way but not the reverse once serialized revisions and life‑limited components are considered. In fashion, explicit and implicit substitutions shift with the merchandising calendar and size-color ladders. In home improvement, items are bound by project‑level dependencies: a faucet without the matching trap kit or connectors is dead stock. In groceries, short shelf‑lives meet aggressive markdown policies, secondary channels, and supplier rebates that reprice inventory within days. E‑commerce couples SKUs at dispatch through packaging, carrier rules, and delivery cut‑offs. Each domain’s “cables” are commonplace inside the trade and fatal to generic templates.
Because supply chains are systems, such quirks do not remain at the margins. They couple to the rest of the flow through shared capacities, baskets, and customer expectations. Ring-fencing them for manual treatment merely shifts cost and risk. If they could be quarantined, a small, semi-autonomous process would do; they cannot. A constraint unearthed in one corner—coil lengths, serial compatibilities, size-color ladders—alters the feasible assortment, reprices options, shifts inbound cadence, and reshapes order assembly and the service promise.
Software vendors face a stark engineering law when they try to absorb these couplings into a packaged product: interactions grow faster than capabilities grow. Each new “core” capability that truly interacts with the others multiplies test surfaces and edge cases; a modest 25 % expansion in such capabilities can easily double the engineering burden. The cost curve is super-linear because the product must cover many firms’ idiosyncrasies at once, whereas each firm only needs its own.
In contrast, the crude programmatic instruments at hand—above all spreadsheets—let practitioners focus on the few complications that actually exist in their context. By narrowing scope to their own economics, they sidestep the diseconomies that sink generic resolutions. This, more than habit, explains why teams fall back to spreadsheets after expensive “planning” deployments.
None of this indicts systems of records. CRUD applications that store authoritative events and enforce rule‑based workflows can be packaged; their entities are largely decoupled by design and their couplings are explicit, local, and auditable. The trouble begins only when the same packaging posture is extended to a system of intelligence, whose very purpose is to navigate firm‑specific couplings under uncertainty.
Packaged decision engines would be a comforting prospect for non-technical managers; they would demote programming to a second-class skill set. Unfortunately, the nature of systems of intelligence rules this out. Short of general artificial intelligence, no finite catalog of templates can anticipate the couplings that matter to a given firm at a given time. The quest for configurability in place of programmability is wishful thinking.
Thus, we are left with the prospect of leveraging ad hoc programmatic resolutions for supply chain. The task is less daunting than it might look because spreadsheets are programmable, and this aspect has not prevented spreadsheets from enjoying considerable success within supply chain circles. However, it raises the question of the adequacy of the tools we have for this undertaking, starting with the choice of programming paradigms. Thousands of programming languages have been devised since the 1950s. It is unreasonable to assume that they are all equally suitable for tackling supply chain challenges. In particular, the popularity a language or tool enjoys beyond supply chain circles has little to do with its suitability for addressing supply chain challenges.
Experimental optimization
Supply chain challenges are at the very least opaque, and occasionally wicked as well. An opaque challenge is characterized by high epistemic uncertainty: we face known unknowns, and often unknown unknowns. A wicked challenge is worse: the challenge shifts as one attempts to address it—even when the actor is a third party, typically a competitor. Wicked challenges do not abide by any static formal resolution.
Opacity is an inherent property of the systems of records that underlie the execution of the flow. As discussed in Chapter Information, merely establishing correct semantics for all the relevant data represents a vast and error-prone undertaking. In practice, even a modest numerical recipe cherry-picks hundreds of fields from among tens of thousands of fields spread over hundreds of tables.
Wickedness arises whenever the adequacy of the response depends on what competitors are doing or not doing. For example, a premium-price, high-availability stance can thrive when most rivals chase low prices and thin stocks; it may falter as soon as they converge on the same premium, high-service posture.
Experimental optimization147 is the cornerstone of numerical engineering for supply chain. It addresses, head-on, both the opacity—and the occasional wickedness—of supply chain challenges. Experimental optimization leverages “real-world” feedback to iteratively improve the numerical recipe. First, it weeds out the mundane coding mistakes of economic import. Those mistakes are both varied and frequent, and cannot be avoided by merely scrutinizing the code. Second, it paves the way for discovering the adequate optimization criteria. Without experimental optimization, attempts at deploying superior software technologies for supply chain invariably fail, and practitioners end up reverting to their spreadsheets.
Cartesianism
Supply chain authors—and, by extension, most enterprise software vendors—treat numerical recipes as a top-down, purely intellectual undertaking. They begin by positing the situation—SKU counts, network topology, constraints, metrics to be optimized, and similar particulars. From this baggage, the expert derives a solution. The solution typically takes the form of a framework for composing the final numerical recipes. The framework is introduced to accommodate variation and to be reusable across many situations.
This approach is known as the Cartesian perspective, named for the 17^th^-century French philosopher René Descartes. Cartesianism places heavy emphasis on pure reason and the power of the mind. It holds that arriving at truth is, at bottom, a strictly intellectual exercise. From a scant set of self-evident truths, careful, methodical reasoning should reach deeper, more complex conclusions.
While reason is certainly preferable148 to authorities, mysticism, corrupt incentives, or drug-induced visions, Cartesianism is not without defects—especially in supply chain. Indeed, nothing is self-evident about the initial baggage that purportedly characterizes a given supply chain situation. In practice, this baggage is a haphazard, inordinate assortment of items of questionable validity.
Authors usually know that a purely intellectual framework is an unsatisfying proposition. Accordingly, they implement the framework and run numerous in silico experiments on it. These take the form of benchmarks, back-tests, and other simulations. These “experiments” are intended to demonstrate the framework’s soundness and performance under plausible conditions.
Unfortunately, subjecting a framework to a “plausible” battery of tests proves little—certainly not its soundness in operating a real-world supply chain. At best, such tests catch gross issues—scalability woes, for instance—that would otherwise impair operations. Yet these in silico validations are nowhere near the proof we seek. We seek strong guarantees that deploying the resulting numerical recipe will be net positive for the company.
Such guarantees come not from added formalism but from falsifiability in the Popperian sense (see Chapter Epistemology). Each numerical recipe is a conjecture about how the firm should commit scarce resources; its premises must be exposed to failure in the world it intends to steer. Benchmarks and back-tests are useful pre-flight checks—they probe internal consistency and scalability—but they do not confront adversarial incentives, semantic drift, and fat-tailed events that govern live flow. The appropriate remedy for the Cartesian posture is, therefore, methodological, not algebraic: engineer the environment so the recipe meets reality early and often, and make refutation the main source of progress.
Mundane issues
When a supply chain practitioner implements a numerical recipe, the first batch of generated decisions is all but guaranteed to be unsatisfactory. Many figures the recipe produces will be “insane”—a term we revisit shortly. The root causes of those insane decisions are usually simple. Five classes commonly undermine numerical recipes: mundane bugs, data semantics, constraints, economic drivers, and strategy.
First, mundane bugs. A numerical recipe worthy of a real-world supply chain spans thousands of lines of code—to say nothing of the millions in generic dependencies. Seven decades of software engineering have taught one lesson: the only code without bugs is the code unwritten. Suitable paradigms may improve matters, but complete elimination is nowhere in sight—especially when combining software written by many hands, where issues emerge at the boundaries.
Second, data semantics woes. This class has already been discussed in Chapter Information. Fundamentally, the semantics of any database column—say, “Order Date”—are in the eye of the operator. If an operator takes “Order Date” to mean the date the supplier acknowledged his willingness to serve the order, then that is the semantics, whatever the system-of-records documentation says. Given that a modern system of records has tens of thousands of columns, many of them interpretable in multiple ways, misunderstandings are unavoidable. Such misinterpretations can severely distort the generated decisions.
Third, constraints are often either untold or invented. An untold constraint is one not properly documented. For example, Bob, who handles warehouse replenishments, knows that above 5,000 inbound units per day the team struggles and must resort to expensive overtime to keep up. Bob learned this a year ago while spending a few months on the warehousing team. Yet this throughput limit was never documented, so a numerical recipe ignores it entirely. Conversely, invented constraints arise whenever the real constraint is deemed too convoluted for semi-manual resolution. For example, price breaks offered by a supplier are often replaced with an MOQ (minimum order quantity) reflecting whatever per-unit price the purchasing team deems acceptable. Yet the constraint is fabricated: it is possible to order below the so-called MOQ, provided the company accepts a higher per-unit price. A recipe that misses this nuance will often overstock by forcibly reaching for the MOQ.
Fourth, economic drivers can be incorrectly assessed. As seen in Chapter Economics, valuation is hard. Analyzing the flow economically requires numerous heuristics for spending and revenue attribution. Moreover, many long-term effects can only be estimated crudely. For example, discounts are likely to erode the customer’s willingness to pay, making him increasingly opportunistic about waiting for the next promotion. Quantifying this effect is challenging, as the relevant time spans are measured in years. A recipe driven by incorrect economics generates decisions with the optics of prudence that, over time, fail to deliver the expected returns.
Fifth, the company’s strategy may look good on paper but not in execution. For example, leadership may prioritize “ultimate customer satisfaction”, only to learn that leniencies—liberal returns or overnight shipments—are abused by bad actors. More generally, strategy usually reflects, to a large degree, an intuitive grasp of the company’s underlying economics. Thus, the understanding can be directionally correct yet miss the fine print, which may demand selective use of alternative strategies. More broadly, the nature of markets—and their unrelenting change—guarantees that, with time, any given strategy will be undermined by competitors.
The odds of obtaining suitable decisions from a newly drafted numerical recipe are vanishingly small. Chasing “First-Time-Right”—a perspective common in manufacturing—is wishful thinking in supply chain. Ambient complexity and opacity are too high for this. It is, however, very much possible to improve this original draft. It boils down to identifying—and then eliminating—the “insane” decisions.
Insane decisions
A supply chain decision is “insane” when it will, most likely, generate unintended losses for the company. A freshly drafted numerical recipe invariably generates numerous insane decisions. Beyond immediate financial damage, insane decisions rightly undermine practitioners’ trust in the automation itself. Left unaddressed, their very presence in the outputs guarantees that practitioners will ditch the recipe and revert to spreadsheets.
Note first that while identifying insane decisions is a problem of general intelligence (cf. Chapter Intelligence), it is not a problem of great intelligence. A careful manual review—one figure at a time—by a practitioner suffices to catch most offenders. The primary hurdle is not the practitioner’s capacity but the sheer time required to perform such a review at scale.
The difficulty is akin to identifying implausible plot twists in a novel. Guidelines may expedite the task, but there is no hope of fully addressing this challenge within the confines of a narrow formalization. Yet most readers can readily tell when a story ventures beyond what can be considered “plausible”.
Counterintuitively, insane decisions are often the only way to surface flaws in the numerical recipe. Confronted with an insane decision, one must reverse-engineer the sequence of calculations that led to it. In software, a defect almost invariably reflects a mistaken belief. Something is believed true when, in fact, it isn’t. By its very existence, the insane decision helps the practitioner pinpoint which belief is incorrect.
Iterative falsification
Experimental optimization is a methodology that seeks to falsify the numerical recipe as it presently stands. It recognizes that supply chain decisions must be subjected to repeated trials against reality, and that these trials play a fundamental role in exposing hidden mistakes in even the most carefully conceived solution. Rather than pretending the problem is fully pinned down in advance, experimental optimization seeks to uncover and fix problems through an iterative, hands-on loop.
The process unfolds as follows:
-
Propose a numerical recipe—a software routine or script—that converts available data (forecasts, costs, constraints) into explicit supply chain decisions (purchase orders, production runs, allocations).
-
Run the recipe in an environment stable enough to generate real or near-real decisions, forcing it to face the raw complexity of a live dataset—with no pretense that “difficult” products or sites can be conveniently excluded.
-
Identify “insane” decisions; this task typically demands human scrutiny: managers or analysts inspect outliers, suspicious allocations, or jarringly large orders until they locate the root cause. The cause may lie in incorrect data semantics, an overlooked cost driver, or a newly emerged constraint.
-
Refine the instrumentation and revise the recipe: missing data or features (e.g., lead-time variance, new categories of supplier constraints) drive the next revision. The explicit cost structure may also be adjusted if the iteration reveals that the penalty for stockouts is unrealistically low or induces wasteful overstock.
-
Repeat until each loop ceases to generate new “insane” results; over time, glaring defects diminish and solutions become more robust and more aligned with the economic reality of the supply chain.
A distinctive strength of this iterative approach is that it invites direct falsification. If the recipe’s proposed decisions look problematic—even for a few SKUs or sites—some assumption must be challenged and corrected. This contrasts with the purely mathematical optimization view, where proofs of optimality offer little recourse when the real world fails to match the original model. Experimental optimization welcomes such mismatches: each defect or “insane” recommendation is an opportunity to expose deeper errors.
Instrumentation
Experimental optimization depends heavily on instrumentation. The data pipeline must allow swift preparation and execution of each new iteration. If each revision takes weeks to deploy, the process cannot keep pace with its environment. Hence the need for carefully versioned data, reproducible workflows, and the ability to “time-travel” by inspecting past states and results.
One might expect metrics—inventory turns, fill rates, forecast accuracy—to be the final arbiters of success or failure. These measures matter, but they lag events and can be warped by changing conditions. Moreover, metrics can be gamed. Companies under pressure to boost service levels, for example, may simply overstock across the board. A few months later, the inflated stock triggers cost overruns. In response, new constraints are layered atop the original system. This back-and-forth between metrics and behaviors can devolve into a tangle of shifting rules that obscure performance.
In experimental optimization, the clearest signal of progress is the absence of glaring missteps. Each day the software’s proposed decisions appear sane without manual fixes or last-minute overrides. If the supply chain teams no longer need to “firefight” for half their working hours, the business is already capturing better performance: fewer disruptions, fewer frantic reshipments, fewer odd workarounds on the warehouse floor. Over time, these operational gains reflect deeper alignment of decisions with correct cost structures and more reliable data semantics—improvements that will eventually show up in standard metrics as well. Even if top-line figures take time to adjust, the disappearance of “insane” decisions is an unmistakable sign of success.
Continuous improvement
Unlike a closed-form approach that purports to “solve” the supply chain once and for all, experimental optimization does not claim convergence to a stable optimum. Supply chains evolve, constraints change, the product mix shifts, new manufacturing plants open, promotions or price wars erupt. Thus the loop never closes. A well-tuned recipe running with minimal insanity today may require substantial updates tomorrow if a global pandemic, a new competitor, or a shift in the firm’s strategic posture arises. Even without external shocks, fresh insights—such as newly discovered cost drivers—warrant further adjustments.
This perpetual flux is no defect. On the contrary, it highlights that supply chain mastery is about orchestrating changes before they cascade out of control. By structuring each change as a small, verifiable experiment, the organization gains both resilience and clarity of purpose. Instead of proclaiming a final formula, supply chain specialists become guardians of a living process—one that recognizes inevitable disruptions as fresh rounds of input.
Experimental optimization merges two ideas that are often kept separate. First, it embraces quantitative ambition: a supply chain is best guided by numbers, systematically handled and recombined by algorithms. Second, it acknowledges falsifiability: the surest way to confirm or refute a method is to subject its output to real conditions. Whether a given plan “makes sense” is known only once the plan confronts raw data on the shop floor, in the warehouse, or at the retail shelf.
By insisting that every proposal runs the gauntlet of reality, experimental optimization transforms errors into catalysts for learning. The cyclical process—instrument, decide, spot insanity, correct, and rerun—does not guarantee that all hidden constraints and cost drivers will be discovered at once. Rather, it ensures that the entire approach, however incomplete or provisional, stays moored to the mission of supply chain: the best possible decisions within irreducibly complex flows. Through repeated trials, the domain’s richness (and unruliness) becomes an engine for incremental progress rather than an insurmountable barrier.
Whiteboxing
An opaque numerical recipe produces outputs—prices, reorder quantities, replenishment plans—yet remains impenetrable to the very practitioners who depend on them. In modern supply chains, any nontrivial data pipeline tends to exhibit this “black box” behavior, in which the numerical recipe involves layers of automation or obscure mathematical models. Even solutions based on straightforward formulas—such as safety stocks derived from “normal demand” assumptions—become black boxes in practice once they handle thousands of items, each with its own nuances and constraints.
Whiteboxing counters this opacity. It recognizes that no advanced data pipeline is intrinsically transparent. Without deliberate effort, the logic of a numerical recipe remains buried within the code, forcing practitioners to choose between two unsatisfactory options: blindly trust the system, or fall back on spreadsheets they can fully control that lack the power or scalability to address the actual complexity of the flow. Whiteboxing provides a third path, preserving strong automation while revealing the “why” behind each recommendation.
A core principle of whiteboxing is that clarity must serve the people on the front line of supply chain decisions. Obfuscation can take many forms—massive data tables, unverified constraints, or advanced machine learning algorithms that only a handful of data scientists can interpret. The response is not to simplify all calculations into toy formulas, which would squander the computational capabilities needed to handle intricate flows. Rather, whiteboxing takes the more challenging route of exposing each decision’s main economic drivers. In addition to raw allocation figures, the recipe should display a short list of maximally relevant economic indicators. For an order recommendation, these indicators could include current inventory, lead time, projected stockout risk (in monetary terms), and the cost of overage if the proposed quantity proves excessive. This handful of columns offers two essential benefits: first, they serve as quick sanity checks for the practitioner; second, they pinpoint where the decision might go awry if any assumption or piece of data proves incorrect.
Such transparency relies on bringing the data pipeline close to the user. Practitioners must be able to add, remove, or adjust the explanatory metrics without launching a months-long IT project. Just as experimental optimization requires rapid iteration to expose and eliminate “insane” decisions, whiteboxing requires rapid iteration in the dashboards themselves. If a decision looks off, the team refines or adds an explanatory indicator, reruns the pipeline with time-travel reproducibility, and studies whether the newly revealed driver explains the decision or exposes a bug. This approach means that every new indicator—be it a lost-sales penalty or a container-fill cost—can be swiftly tested and adjusted. Over time, these indicators cohere into a compact suite of cost and risk measures that shed light on each recommendation without inundating users with trivial metrics.
Crucially, whiteboxing guarantees neither universal consensus nor perfect modeling of all constraints. Instead, it seeks to ensure that decisions do not trigger abrupt disbelief or “panic overrides” by practitioners who cannot see the logic behind them. Transparency lowers the threshold of trust needed to adopt automation. When supply chain teams can tie a proposed production run or reorder quantity to a visible trade-off—say, the cost of a potential line stoppage versus the holding cost of extra parts—they more readily accept the numeric result, even if it contrasts with past rules of thumb. Conversely, when an outcome defies intuitive checks, whiteboxing directs attention to which driver may be missing or mislabeled, guiding the next improvement.
Another vital dimension of whiteboxing is its capacity to reduce organizational friction. In many companies, forecasts or replenishment plans move from data science teams to supply chain managers, then to finance or warehouse operations, each group harboring its own metrics, data definitions, and constraints. This complex handover often multiplies confusion: a manager seeing “buy 1,200 units” in one system might not know that a quantity-based discount was factored in elsewhere. By revealing the principal costs and constraints directly in the final decisions, whiteboxing bridges these handoffs. Stakeholders can see, in monetary terms, how each driver was weighed. They can correct or refine those values without dismantling the entire pipeline.
Far from a one-off task, whiteboxing serves as an ongoing design principle. As soon as a supply chain evolves—through new suppliers, changed lead times, or adjusted strategic goals—so must the transparency mechanisms. If not, hidden or outdated assumptions resurface, creating new black-box zones. Sustaining whiteboxing demands regular collaboration between domain experts and the teams responsible for engineering the numerical recipes, ensuring that whenever an aspect of the flow changes, the explanatory metrics change as well.
Whiteboxing thus stands at the intersection of reliability and credibility. It addresses the practical realities of modern supply chains, where advanced numerical methods are indispensable, yet so is the trust of the people who put those methods to work. Without whiteboxing, the best-engineered system remains unusable; with it, even sophisticated processes become comprehensible enough for practitioners to adopt, refine, and operate day after day.
Fermi-ization of the elusive
No matter how one approaches the numerical recipe, supply chain teams inevitably encounter elusive parameters. These are quantities or forces that do not readily appear in any system of record and thus elude direct measurement. They can be strategic—for example, the expected release date of a promising, still-in-development technology on which a product roadmap implicitly depends. Or they can be operational—such as the probability that a large but financially stressed supplier will file for bankruptcy within the next few quarters. When confronted with such factors, the traditional approach is to bypass the problem. Practitioners might leave these parameters out of the automation altogether and revert to a semi-manual process guided by a mixture of opinion, guesswork, and cautionary hedges. That workaround dooms the numerical recipe to dependence on continuous human intervention, which undermines any prospect of genuine automation.
A better approach treats these “unmeasurables” as prime targets for a Fermi-style analysis, akin to what Philip E. Tetlock and Dan Gardner describe in “Superforecasting: The Art and Science of Prediction” (2015). The key insight, popularized by the physicist Enrico Fermi, is that a seemingly impossible quantity can be broken down into a series of more tractable estimates that, when multiplied or added, produce a credible estimate of the original unknown. Fermi himself famously asked students to estimate how many piano tuners worked in Chicago—a figure most would deem unknowable absent an official listing. Yet by dividing the problem into small, separate pieces (such as the approximate population of the city, the fraction of households with pianos, and how often a piano gets tuned), the final estimate can land surprisingly close to the truth.
In supply-chain practice, a more telling illustration is the probability that import tariffs on goods from Ruritania will double over the next five years. This quantity appears in no ledger, yet it directly governs buying cadence, sourcing mix, and pricing. A Fermi decomposition renders the estimate tractable. Start with an outside view: tally, over the last century of Westonia–Ruritania relations, how often tariffs of comparable magnitude were raised or lowered, and over what horizons. This historical base rate anchors the prior firmly. Then incorporate the inside view: read the present political discourse in Westonia; judge whether the party in power argues for higher or lower tariffs and whether the topic is salient or peripheral to its platform, and adjust the prior accordingly. Next, weigh plausible election outcomes over the relevant horizon—say, two years ahead—using polling or market odds. Conditional on each outcome, assess the likelihood that the winning party would pursue a doubling. Finally, consider retaliation pathways: a tariff initiated by Ruritania may trigger reciprocal measures by Westonia. Mirror the same reasoning from the Ruritanian side—its domestic politics, its historical propensity to use tariffs, and the salience of trade tensions—to price the chance that Westonia’s move is reactive rather than proactive. The end product is not a point estimate but a calibrated probability that can be revised as information accrues.
Once this probability exists, it can be translated into money-valued decisions: expected duty uplift on affected SKUs; the option value of accelerating or deferring imports ahead of cutoffs; the premium justified for alternative origins with lower exposure; and the merit of postponement (late binding) to keep tariff-sensitive transformations offshore. The estimate can be rough; what matters is that it is explicit, auditable, and placed on the same footing as purchase and holding costs, enabling unattended software to weigh it alongside other drivers.
Crucially, this decomposition of the elusive parameter should not stop once the initial guess has been coded. Under experimental optimization, these newly introduced parameters remain subject to the same falsification loops as everything else. The fact that “tariff-doubling probability” was not in the system of records does not exempt it from scrutiny; on the contrary, its novelty calls for vigilant watchfulness. If an insane decision emerges—for instance, one that systematically rushes oversized inbound orders and starves working capital to preempt a shock that never materializes—this signals that the probability (or its economic translation) is inflated. Conversely, if duties do jump and the numerical recipe shows no sign of prior hedging—no pulled-forward shipments, no origin diversification, no postponement—then the probability was understated. Refinement is iterative: each pass, guided by real outcomes or near misses, tightens the logic until no obviously insane decisions arise.
In practice, the Fermi approach also becomes a powerful guardrail against the cognitive biases that attend any guess. These guesses are not conjured from a vague impression but anchored in a chain of smaller, more tangible estimates. The process fosters “active open-mindedness”, a trait documented among Tetlock’s superforecasters, who methodically break down a problem and then weigh outside perspectives to calibrate each piece. The entire supply chain team can see—directly in the code or in a white-boxed dashboard—how this tariff-doubling probability was computed, which fosters alignment and invites collaborative fine-tuning rather than unilateral overrides. If an import manager judges that the election weights, or the salience of trade policy in a manifesto, is off, the team can revise that single sub-estimate rather than tear apart the method.
Although it is laborious, Fermi-ization is a straightforward extension of the techniques already needed to keep a numerical recipe fully automated. Once supply chain practitioners accept that no parameter is too intangible to be broken down—even if it must be done with rough heuristics—the path to including those elusive elements in the system becomes clear. Just as with “normal” parameters, the guess is best introduced under rigorous instrumentation, so that each production run can confirm or challenge the initial logic. Supply chain mastery thrives on iteration rather than one-off determinations. When confronted with the intangible, the biggest mistake is not misestimating initially; it is refusing to produce any estimate at all, thus dooming the recipe to partial manual operation.
By fully embracing the Fermi principle, supply chain practitioners can tackle exactly those domains previously deemed “beyond measurement”149. Instead of bowing to intangible elements that incapacitate automation, they transform them into a set of smaller, more tractable pieces that can be tested and refined. As with all parts of a living numerical recipe, each approximation is linked to the real-world results it produces, allowing supply chain to integrate what it learns over time. The alternative—silently discarding such elusive parameters as hopeless—ensures that automation remains little more than an aspiration. Fermi-ization offers a resolutely practical path to making even the most elusive parameters explicit, ensuring that all critical drivers of the flow, intangible or not, can be folded into a holistic automated system.
On programmability
In previous chapters, we argued that modern supply chains must be steered by unattended software that issues auditable, economically priced commitments. Once we accept a programmatic resolution to supply-chain problems, a question that is mundane in appearance but foundational in fact arises: in which language should these decisions be expressed? This is no tooling whim. A language fixes which ideas can be made explicit, which options can be enumerated without contortion, how uncertainty is priced, and how resulting commitments are audited years later.
The field has mostly dodged this question since the end of the 4GL episode.150 The vacuum has been filled by spreadsheets, rule engines glued to ERPs, and scripting in general‑purpose languages. Each of these instruments has its place, but none is adequate to serve as the native medium for unattended decision-making at supply-chain scale.
Systems of records speak SQL because they must preserve facts, not invent them. SQL excels at stating what is inside a ledger—rows and their relations—not what should be done next in the face of variability and trade‑offs. Rule engines mounted on top of transactional systems try to bridge the gap with a drip of if‑then clauses; they age into brittle accretions that no one fully understands, least of all when incentives change. Scripting languages—today most often Python, yesterday VBA151, earlier still proprietary macros—are powerful prostheses for analysts, but their very freedom undermines the properties we need most in production: determinism, reproducibility, and strict control of side effects. They produce artisanal tooling; unattended decision engines must be industrial.
A suitable language for supply-chain decision automation must align with flow economics and the physics of computers. Its objects should be the ones practitioners actually manipulate: typed tables keyed by SKUs, locations, and dates; arrays over those tables; and—crucially—random variables that carry demand, lead time, and return distributions. Uncertainty must be first-class, together with an algebra that converts distributions into currency-denominated payoffs so that the expected, risk-adjusted return of any move can be written as a short formula rather than an after-the-fact simulation. Prices must be explicit: currency conversions and rounding as built-in functions. Time must also be first-class: calendars and calendar algebra. When these primitives exist, the engine expresses the decision directly; when they are missing, practitioners smuggle them in piecemeal and the codebase degenerates into workarounds.
Because unattended engines live or die by auditability, execution must be deterministic. The same inputs must yield the same outputs—down to the last cent—even when computation is parallelized. Randomness, when needed for stochastic simulation, must be explicit and reproducible. Side effects must be quarantined at the boundary: read authoritative records, compute decisions in a pure core, then write the resulting commitments back to the ledger. This separation of concerns is not academic tidiness; it is the only way to make ex post audits and dual‑run comparisons meaningful.
Decision logic over millions of SKUs, sites, and dates cannot be written as a procession of row-by-row loops without buckling under its own weight. The language therefore needs array or table semantics—vectorized operations over large, typed collections—so most programs read as algebra over sets rather than choreography over iterators. Such designs encourage mechanical sympathy: they map well to caches and parallel units and make large problems cheap enough to solve often.
The question declarative or imperative? is ill-posed in the abstract. We need an imperative shell to stage the flow—ingest, compute, emit—and a mostly declarative core to express valuations and constraints as formulas over tables and distributions. In practice, this means feasibility tests, penalties, and prices are written as expressions the engine can analyze, reorder, and partially recompute. The resulting transparency enables halting heuristics: when confidence drops, the engine stops on its own terms instead of pestering humans with exceptions it cannot resolve.
None of the above is about syntactic taste. Braces versus indentation is trivial. What matters is semantics: first-class uncertainty, money and time as citizens rather than afterthoughts, side-effect isolation, determinism, vectorized evaluation, and a runtime that produces dossiers fit for audit. General-purpose languages can be bent toward these aims via conventions and discipline; experience shows that at scale, and over time, conventions decay. A supply‑chain language worthy of the name bakes these properties into its very grammar.
The remainder of this section operationalizes the selection criteria. We examine how language semantics bear on unattended execution (determinism and reproducibility), on economics (native money, prices, and penalties), on uncertainty (probabilistic primitives and stochastic simulation), and on maintainability (small cores, legible programs, diff-friendly artifacts). The goal is simple: choose a language that lets the firm write its economics down once, run it at machine speed, and defend every commitment ex post in coins.
Programming paradigms
A programming paradigm is the framework that shapes how software is conceptualized, structured, and executed. It fixes the principles for organizing data, defining operations, and determining control flow, guiding both design and execution.
Developing suitable programming paradigms for supply chain is arguably the single most important technical contribution to the field. Indeed, a suitable programming paradigm facilitates the resolution of most—if not all—supply-chain challenges, irrespective of their specific complications. In this regard, programming paradigms are wholly dissimilar to “models” or any other kind of closed analytical solutions, which are, by design, brittle when confronted with unplanned complications. While many techniques are of broad interest—stochastic gradient descent would be a perfect example—programming paradigms are unrivaled in this regard. The short list below presents some of the most relevant programming paradigms for supply chain.
Vertical integration names an environment that provides the entire stack—from data storage, through processing and logic, to presentation—inside a single, coherent medium. The canonical exemplar is the spreadsheet: a monolithic application where the grid is at once the datastore, the formula engine, and the user interface, with live recalculation binding the three. In such an environment there is no impedance mismatch between “where the facts live”, “how they are transformed”, and “how they are inspected”; one artifact carries them all.
Vertical integration matters because it collapses handoffs. It lets a small number of practitioners who are not software specialists author, run, and audit numerical recipes end-to-end. The same person can ingest records, express the decision logic, and visualize the resulting commitments without waiting for another team to expose an API or wire a dashboard. Absent vertical integration, composing the recipe becomes a technical project that cannot be left to anyone but software experts. The effort metastasizes into the usual procession—supply-chain experts, consultants, project managers, database administrators, software engineers, data analysts—whose coordination creates an interface-design problem larger than the original economic problem. Spreadsheets demonstrate both the attraction and the limits of this pattern: they make vertical integration immediately tangible, yet they also reveal why, beyond a trivial scale, a different substrate is required (see the section “Spreadsheets” below).
Time-travel reproducibility means that the environment preserves the exact code and data as they existed at any point in history. Users can step back in time and rerun the computations as they were. This type of end-to-end versioning (code + data) ensures that results are not only reproducible in principle but also automatically verifiable at any historical point.
Indeed, in supply-chain decision-making, even small glitches can prove extremely costly if bad decisions are generated152. Too often the problematic behavior cannot be reproduced because, by the time the issue is identified, the input data has changed. As a result, glitches remain unaddressed far longer than desirable. Time-travel reproducibility is critical for eliminating those defects before supply chain teams lose faith in the viability of the numerical recipe and revert to their spreadsheets.
Relational algebra operates over tabular data—typically stored across multiple tables—through a concise set of composable operations such as selection, projection, join, and aggregation. These operations, formalized in the 1970s,153 gave rise to the SQL (Structured Query Language) family of technologies that remains ubiquitous in modern enterprise systems. The chief virtue of relational algebra is its declarative nature: rather than spelling out how to perform operations step by step, developers and analysts specify what results they want, and let the execution engine handle the mechanics.
Relational algebra is paramount for supply chain because the overwhelming majority of the operational data—sales transactions, purchase orders, stock levels, transportation events—resides in relational databases. The relational model offers a systematic way to link these pieces, ensuring that joins on common keys yield a coherent whole. Without relational algebra to elegantly unify these disparate sources, practitioners fall back on a tangle of ad hoc workarounds—typically unwieldy macros in spreadsheets—which are not only error-prone but also computationally slow. In practice, a relational algebra is indispensable for any numerical recipe supporting a modern supply chain, owing to supply chain’s inherent complexity.
Array programming is a paradigm in which operations apply uniformly to entire collections of data—known as arrays—rather than requiring developers to write loops over individual elements154. By treating data in aggregate, array languages and libraries155 enable concise expressions that mirror the mathematical formulation of the problem. Instead of writing error-prone loops, the practitioner specifies higher-level transformations—such as arithmetic, logical, or statistical operations—that automatically broadcast over each element of the array. By eliminating the need for explicit indexing, array programming significantly reduces off-by-one errors and other subtle defects. Moreover, because the underlying execution engines can optimize bulk operations, array code frequently outperforms equivalent loop-based routines, particularly on large datasets.
From a supply chain perspective, array programming is critical because nearly all relevant decisions must be computed for thousands (if not millions) of SKUs, store locations, or time steps at once. Iterative or loop-based approaches make this process slower to write, harder to debug, and more prone to silent mistakes, especially when coping with edge cases or intricate constraints. In contrast, array programming’s vectorized style yields compact, expressive code that is not only easier to review but also simpler to extend as new product lines, locations, or supply chain complexities emerge. This heightened clarity reduces the risk of subtle logic bugs going undetected, while the performance gains of vectorized operations enable faster iteration. In practice, these two attributes—correctness and agility—greatly improve a company’s ability to prototype, refine, and confidently deploy its decision-making processes.
Distributed dataflow is a programming paradigm that automatically distributes computations across multiple machines, effectively partitioning and parallelizing workloads to accommodate vast volumes of data. Rather than forcing developers to orchestrate concurrency on a step-by-step basis, distributed dataflow systems typically represent data transformations as Directed Acyclic Graphs (DAGs), which can then be executed in parallel by a cluster of machines156. This approach promotes resilience to partial failures—faulty nodes can be bypassed or replaced without halting the entire job—and naturally leverages modern cloud infrastructures. By abstracting away most low-level details of network communication and process synchronization, distributed dataflow enables practitioners to focus on the logical flow of their calculations instead of wrestling with concurrency primitives.
For supply chain, the significance of distributed dataflow is twofold. First, the sheer scale of many operations—spanning SKUs, warehouses, and transportation routes—requires ingesting and transforming massive datasets. A centralized or sequential system would be prohibitively slow, leaving decision-makers with stale outputs or forcing them to limit the scope of their analyses. Second, it facilitates pooling of computing resources that are only intermittently needed. For example, if a forecast need only be refreshed once per day, the corresponding computing resources can be acquired and released on demand, vastly lowering infrastructure costs. More generally, distributed dataflows are critical to maintaining stable performance under highly variable workloads.
Differential execution is a programming paradigm in which the system automatically tracks dependencies between calculations and re-executes only those portions of the workflow affected by a given change. Rather than re-running the entire pipeline from scratch, the engine inspects a graph of computation—representing data transformations as nodes and edges—and evaluates a “diff” between the current graph and the one previously executed. Intermediate outputs that remain valid are preserved, while only the new or modified nodes are recalculated. Additionally, the system can intelligently cache expensive or time-consuming intermediate results that yield relatively small outputs. By doing so, it accelerates subsequent runs that involve the same partial computations. This paradigm stands apart from both standard batch processing and purely incremental approaches, as it combines the breadth of a full dataflow engine with the selectivity of fine-grained updates.
From a supply chain perspective, differential execution is critical because numerical recipes are inherently iterative. Practitioners typically edit a few lines and then rerun the recipe—sanity-checking the results computed so far. Given that the final recipe often spans thousands of lines, its development involves hundreds of small iterations. Without differential execution, each small tweak to the script would require reprocessing all datasets in full. This is prohibitively expensive and discourages rapid iteration, especially when dealing with gigabyte-sized datasets flowing in from enterprise systems. By zeroing in on the minimal set of computations that need refreshing, differential execution preserves resources, reduces turnaround time, and fosters fast feedback loops.
Algebra of random variables is a programming paradigm where uncertain quantities are modeled and manipulated as random variables—not merely as numerical scalars or point estimates. In this environment, addition, multiplication, and other arithmetic or logical operations can be applied directly to random variables, yielding new random variables. Rather than bundling all uncertainty into error-prone ad hoc simulations or unstructured data tables, practitioners can define each uncertain factor—demand, lead time, buy prices, production yields, or returns—as its own distribution. When these factors interact, the language seamlessly combines their distributions through a coherent set of probabilistic operations. The resulting code closely mirrors the underlying stochastic phenomena, bypassing complex looping or separate Monte Carlo steps. By treating random variables as first-class citizens, this paradigm enables precise and transparent transformations of uncertainty throughout the entire decision-making pipeline.
This algebraic treatment of uncertainty proves vital for supply chain, given the pervasive role of probabilistic forecasts in risk analysis. In real-world operations, it is never sufficient to track a single best guess: demand often spikes without warning, parts arrive late for reasons no one fully understands, and exchange rates fluctuate in ways that mock simple predictions. Without an explicit “algebra” to combine these distributions, simplistic point-based analytics are almost invariably used, which then silently discard information as soon as averages or worst-case assumptions are taken. By contrast, an algebra of random variables preserves the entire probability distribution from upstream (e.g., supplier lead time) to downstream (e.g., inventory placement), ensuring that uncertainties remain visible, consistent, and mathematically grounded. This approach not only reduces the likelihood of costly miscalculations—where a small underestimation can lead to stockouts or a slight overestimation triggers inflated holding costs—but also accelerates experimentation.
Differentiable programming is a paradigm that fuses the flexibility of general-purpose programming with the computational techniques of automatic differentiation and gradient-based optimization. At its core, this approach treats every step in a data transformation pipeline as a function amenable to differentiation, allowing parameters to be tuned via stochastic gradient descent (or related optimizers) directly within the code that implements the logic. Rather than relegating learning to an isolated environment (e.g., exporting data to a specialized machine-learning toolkit), differentiable programming makes “trainable code” a first-class citizen. Each transformation—from input parsing to final forecast—can be designed, annotated with parameters, and automatically updated to minimize a chosen loss function. This paradigm goes beyond purely numerical procedures: it weaves domain insights straight into the parametric structure of the models, ensuring that these structures remain interpretable and aligned with the underlying business processes.
In supply chain, differentiable programming is particularly potent because it bridges several critical gaps simultaneously. First, supply chain data is notoriously sparse and heterogeneous; many SKUs, locations, or time segments offer only limited observations, yet share higher-level patterns (e.g., weekly cycles, promotional lifts). A differentiable program can explicitly model these patterns through shared parameters—thereby improving statistical efficiency compared to off-the-shelf black-box approaches. Second, the very shape of the supply chain problem—its mix of relational queries, array transformations, and random variables—can be encoded directly in a parametric, learnable script, rather than forcing practitioners to patch together a “language-in-language” design. This alignment not only streamlines iteration but also yields more transparent models whose named parameters (e.g., “seasonal amplitude”, “port congestion shift”, “Black-Friday uplift”) resonate with domain experts. Above all, by integrating learning loops into the normal flow of data and decisions, differentiable programming offers a unified environment in which forecasting and optimization can co-evolve with the actual business constraints—rather than stand apart as a one-off or purely academic exercise.
Deterministic Distributed RNG is a programming paradigm designed to reconcile two seemingly conflicting goals: leveraging large-scale Monte Carlo simulations across multiple machines while ensuring that every run—even across a distributed cluster—produces bit-for-bit identical results given the same inputs. Put differently, the approach guarantees that random draws, once seeded, unfold in a predictable pattern throughout the dataflow graph, irrespective of shifting workloads or dynamic machine assignments. Monte Carlo techniques often surpass a direct algebra of random variables when the phenomenon in question resists analytical decomposition; however, naive parallelization can easily render the process non-deterministic, making it impossible to reproduce outcomes precisely from one run to the next. Deterministic distributed RNG averts this pitfall through a carefully orchestrated mechanism of pseudo-random seed management and partitioning, thereby combining the flexibility of simulation-based approaches with the reproducibility vital for rigorous engineering.
This paradigm is indispensable in supply chain for two main reasons. First, the domain is already rife with genuine uncertainty: demand shocks, supplier delays, currency swings, and more. If the software layer itself introduces its own randomness—generating different outputs on each re-run despite unchanged inputs—managers cannot reliably diagnose issues or trust the decision-making pipeline. Second, supply chain initiatives live or die by continuous iteration, as practitioners frequently refine their numerical recipes, replay historical scenarios, and compare new results against old baselines. Strict determinism spares them from chasing “heisenbugs” that vanish when retested and ensures that any anomaly discovered a week or a year later can be traced to the precise combination of data and code that caused it. By preserving exact reproducibility in their simulations and forecasts, supply chain teams gain auditability and confidence needed to safely automate decisions.
Spreadsheets
The enduring success of spreadsheets in supply chain is no coincidence. Spreadsheets offer, within a single grid, a form of functional reactive programming tightly coupled with vertical integration: data storage, formula-driven logic, and a visual user interface coexist in a unified environment. The result is a near-perfect feedback loop—akin to a rapid REPL157—that fosters a gentle learning curve. Supply chain practitioners can start with a few cells, tweak a formula, and instantly see the outcome of their adjustments. This immediacy contrasts sharply with classical software engineering workflows, which require code to be written, packaged, and then deployed before any results can be observed.
Spreadsheets’ programmatic versatility should not be underestimated. Within a matter of minutes, a user can assemble ad hoc logic for forecasting, inventory consolidation, or what-if analyses. The barrier to entry is low; supply chain managers do not need to become database administrators or software engineers to handle many day-to-day calculations. Even as vendors have repeatedly attempted to encapsulate forecasting or optimization in “ready-to-use” analytical modules, spreadsheets have consistently remained popular. This phenomenon arises because the spreadsheet’s functional reactive grid can accommodate edge cases and “cable-like” complications158 more flexibly than rigid off-the-shelf solutions.
Despite these strengths, spreadsheets are fundamentally at odds with the true automation of supply chain decision-making. Practitioners often cite a “lack of scalability” as the reason spreadsheets fail beyond a certain data volume—especially when a workbook is shared via local hosting on a single machine. Although hosting spreadsheets remotely can alleviate some performance burdens, it does not address the deeper, structural limitations. Two such limitations, in particular, thwart the prospect of robust supply chain automation.
First, over time, spreadsheets become riddled with extensively duplicated logic scattered across dozens—sometimes hundreds—of tabs and workbooks. Every time a user copies a formula or re-creates a similar worksheet for a slightly different SKU or location, the structure diverges a little more. Eventually, no one can confidently trace which cells drive the final outcome. Even minor modifications become hazardous, potentially breaking dependencies hidden in cell references. Once multiple people collaborate, these problems multiply. Worse, spreadsheets typically lack the robust versioning, modularity, and auditing features needed to keep track of a fast-evolving codebase.
Second, supply chain intricacies—multi-sourcing, container-filling constraints, lead-time variability, partial bills of materials, multi-echelon hierarchies—rarely flatten into neat two-dimensional grids. Spreadsheets forcibly “slice” each dimension down to rows and columns, complicating even moderate joins or cross-references. For instance, a single SKU that depends on two suppliers with different cost structures and lead times may require contrived cell expansions, plus numerous “auxiliary” tabs that try to simulate a relational model. The more a supply chain requires branching logic or multi-table queries, the more unwieldy the spreadsheet becomes.
Beyond these immediate barriers, spreadsheets also fare poorly against the programming paradigms most conducive to modern supply chain engineering:
-
Time-travel reproducibility becomes nearly impossible with large spreadsheets: while file versions can be saved, each user action modifies the shared document, often without precise control over which cells changed and why. Debugging “heisenbugs” after the fact is nearly hopeless.
-
Algebra of random variables, differentiable programming, or deterministic distributed RNG cannot be practically embedded in typical grid formulas. At best, practitioners tack on external macros or third-party add-ins, but these hacks seldom achieve robust, production-grade workflows.
-
LLM instrumentation—allowing Large Language Models to generate or refactor code automatically—falters with the inherent two-dimensional logic scatter. Most LLM-based approaches assume a more structured textual environment where logical relationships are explicit, not buried in cell references and ephemeral macros.
In principle, spreadsheets could attempt to address some of these shortcomings—for instance, by layering more powerful computational engines underneath the grid. However, the spreadsheet paradigm itself remains unaltered: it is designed for human interactivity and ad hoc exploration, not for systematically handling the ever-shifting complexities of large-scale flows. The moment a company commits to continuous, unattended decision-making—over thousands or millions of SKUs, across hundreds of locations—the cracks show. The more these cracks are patched by short-term fixes, the more convoluted the workbook becomes, until no one can trust the fragile edifice it is.
As a result, spreadsheets, while undeniably central to supply chain practice for decades, represent a technological dead-end. Their downfall is not merely a matter of “more software” or “faster hardware”. Rather, their entire design—grid-based formulas, manual duplication of logic, minimal auditing support, and an implicit focus on local user edits—runs counter to the paradigms needed for large-scale automation. Precisely because spreadsheets excel at one-off or small-scope tasks, they fail to pave the way toward robust, fully automated supply chain processes.
Generic languages
For all their limits, spreadsheets point to a broader truth: supply chain requires ad hoc programming that no static, “packaged” system can satisfy. In principle, any general-purpose (“generic”) language could meet the need, especially now that modern languages are largely multi-paradigm. By stacking libraries, a team might reproduce the essentials for supply chain: time-travel reproducibility, array programming, differential execution, and more. Yet in practice, the path consumes more effort than it returns.
First, generic languages breed accidental complexity. Object-oriented patterns, for example, become needlessly elaborate when modeling mostly tabular supply-chain data. Left unrestricted, network or system calls invite security issues and so-called “supply-chain attacks”—infiltrations not in the company’s code but in a third-party package it depends on. Even purely mathematical operations fracture into a thicket of tiny modules. The deeper a team sinks into this ecosystem, the more time it spends on intricacies that have little to do with allocating scarce supply-chain resources.
Second, relying on a generic environment effectively commits the company to a full-scale software project. The storage layer must be selected or built and integrated; the compute layer must scale for large volumes or SKUs; the presentation layer must be coded or adapted for supply-chain audiences. This sequence of choices, each anchored to a different technology, makes the solution harder to maintain. Any small shift—a new database version, an unmaintained library—can cascade into incompatibilities. Over a few years, “software rot” sets in and siphons resources from actual supply-chain improvements.
Third, the overhead in such a project almost always fragments expertise. A small cadre of software specialists—typically with little supply-chain background—ends up implementing and maintaining the code. Meanwhile, the actual practitioners—those who grasp the knotty quirks and countless other domain-specific complications—are sidelined. They must propose changes indirectly, submitting “tickets” and routing requests through layers of mediation. Agility evaporates as the lag from discovery to deployment widens. In extremis, the solution grows so specialized that only its original authors can patch or extend it—a brittle arrangement when personnel changes or reorganizations occur.
Fourth, while generic languages can in theory support advanced paradigms such as differentiable programming or deterministic Monte Carlo, they seldom do so in a cohesive fashion. The relevant libraries effectively become “languages within the language”. One library manages distributed dataflows; another handles random variables; yet another attempts time-travel data versioning. Each of these subsystems demands its own configuration, data structures, and usage patterns, compounding the cognitive overhead. What was meant as a single decision-making pipeline splinters into separate toolchains, any of which can split at the seams.
None of these challenges is insurmountable: with enough talent, money, and time, the right team can craft a workable solution atop a generic language. Some of the largest retailers and manufacturers do exactly that. Yet this route is so resource-intensive that only firms with tremendous scale can sustain it, and even then it remains fragile. Enhancements in large language models may automate fragments of these workflows—for instance, by generating scaffolding or refactoring older libraries—but they do not resolve the mismatch between a free-form language and the specialized needs of supply chain. Paradoxically, better LLMs strengthen the case for domain-specific environments: the more adept they become at writing and reviewing code, the more a streamlined, supply-chain-oriented language benefits from their assistance.
In short, no absolute technical barrier prevents a generic language from powering a large-scale, fully automated supply-chain process. Yet mounting an initiative this way imposes overhead and complexity ill-suited to most organizations. Because supply chain exists to harness optionality rather than drown in it, layering multiple paradigms and libraries—none of which natively address the field’s pressing needs—usually proves a detour, not a destination.
Deployment
Supply-chain deployment is when the previous chapters—economics, information, intelligence, engineering, and decisions—become a machine that makes sound choices daily. The machine is a system of intelligence that converts authoritative records into unattended, auditable flow commitments. The objective is neither “better visibility” nor a nicer forecast: it is to issue purchase orders, transfers, schedules, and price moves that are economically sound by construction, and to improve them in place.
The deliverable is operational: a working deployment ingests raw extracts from systems of record, carries a money-denominated objective under probabilistic uncertainty, and writes decisions back into the ledgers. Its presence is marked by emissions—lines created in purchase-order tables, stock transfers booked, prices posted—without a planner shepherding each line through a spreadsheet. Anything that stops short of unattended emissions constitutes project theater.
Three properties make this system safe and productive in practice. First, ownership: one small team—ideally a named individual—holds end-to-end responsibility for the numerical recipe, the code that frames options, prices trade-offs, and produces daily commitments. Second, transparency: every emission comes with a dossier that exposes the main economic drivers, so practitioners can see why the decision is sane and finance can audit it in coins. Third, reversibility: the engine is stateless and guarded by halting heuristics; it can be dual-run, promoted, or rolled back within hours, and it halts when confidence drops.
The path to that state is not mysterious, but it is unforgiving. It begins with raw daily extracts—unfiltered, schema-faithful, and stable; continues with a first recipe that runs daily next to the manual routine until no insane lines remain; and culminates in a progressive rollout that retires spreadsheets rather than wrapping the engine in them. Scope is set to capture flows that compete for the same scarce resources; roles are trimmed to what the work requires; governance is economic, not ritual.
Most of what derails deployments is already familiar: pilots kept in read-only mode that never write even a single flow commitment back into the ledgers, partial datasets that hide scale effects, cargo-cult parameters copied from legacy tools, committees that debate forecasts instead of decisions, and procurement rituals that privilege the vendor best at paperwork over the one best at decisions. Avoiding these traps requires the same stance advocated throughout this book: frame the problem in money, surface options explicitly, carry uncertainty faithfully, and let unattended software do the clerical work.
This chapter is a manual for that stance in production: it shows how to pick a partner when one is needed; how to frame scope so the engine touches real constraints quickly; how to structure the data pipeline; how to write, instrument, and harden the recipe; how to dual-run and promote to unattended; and how to keep the system healthy once live, while roles and incentives adjust. The aim is not to add another reporting layer, but to replace manual routines with a capital asset that issues, audits, and improves decisions at machine pace.
Because many firms will not build everything in-house, we begin with the practicalities of selecting a vendor. The method recommended here—adversarial market research—fits the opacity and incentives of software better than RFP theater, and it aligns with the economics-first posture defended in earlier chapters. Once the partner is chosen, we turn to setup and scope, the numerical recipe, system effects, roles, timeline, and the mechanics of rollout and maintenance.
For brevity, the following terminology is used consistently. A system of intelligence denotes the deployable asset that converts authoritative records (plus a few vetted external feeds) into unattended, auditable flow commitments written back to the ledgers. Its logical core—the program that frames options, prices trade-offs, carries uncertainty, and emits lines—is the numerical recipe.
Selecting a vendor
The complexity of modern supply chains far exceeds what can be done without resorting to software. Most lack the means or the inclination to develop those instruments in-house—even if the open-source landscape greatly diminishes the financial burden of such an undertaking. Thus, supply-chain software vendors abound, and most companies, past a certain size, regularly face the prospect of selecting one.
Yet the software industry exhibits extraordinary variety while remaining relatively opaque. The opacity is multidimensional: capabilities, performance, licensing fees, operating costs, reliability, maintainability, evolvability, security, and more.
The open-source landscape should, in principle, be more transparent, since the source code can be freely inspected, but in practice there are two major caveats. First, the technical expertise to reliably assess a given open-source technology is seldom found. Assuming that an open-source technology is sound only because it originates from a well-known source—e.g., a big software vendor—is a time-tested recipe for disaster. Vendors abandon open-source projects all the time because they have little to lose. Second, the open-source landscape deals chiefly in components rather than solutions—especially in enterprise software. This leaves the company, if it wants to leverage those components, rolling out what amounts to an in-house software project.
Thus, for better or worse, companies end up picking software vendors under high uncertainty. Yet software, like most intellectual undertakings, exhibits a wildly uneven range of contributions, from abysmally dumb to peak human ingenuity. Combine the inherent opacity of software with a vendor gamut that includes a thick “dangerously—if charmingly—incompetent” segment, and it becomes imperative for companies to pay close attention to the vendors they pick.
To their credit, most large companies now acknowledge the problem and spend considerable effort on market research for software vendors. No other area attracts as much market-research spend as software. By way of anecdotal evidence, a casual observation of market research firms reveals that software verticals are literally dwarfing all other verticals put together.
Yes, most software market-research efforts borrow methods from other verticals; those methods are dysfunctional for software. Case studies and Request For Proposals are the chief offenders. Software requires a different approach: adversarial market research.
The pitfalls of case studies—often presented as evidence in vendor selection—were examined in Chapter Epistemology. This section focuses on RFPs and a more reliable alternative: adversarial market research.
Request For Proposals
Request For Proposals159 (RFPs) are a bureaucratic power play within large organizations. For software, RFPs are neither rational nor truthful nor informative. RFPs epitomize the bureaucratic mindset: fixation on process at the expense of outcome. In systems of intelligence, the defects are fatal: checklists cannot certify unattended decision quality, halting behavior, or rewritability under pressure. An RFP steers the company toward incompetence. RFPs are too often treated as an inevitable plague in any organization past a certain size. Yet nothing about this self-inflicted plague is inevitable. Eliminating RFPs requires only modest fortitude from top management.
Consider a company whose management faces a software challenge—or opportunity—beyond internal resources. After a few informal exchanges with one or two vendors, management decides a third party will address the challenge.
The RFP mythos purportedly unfolds in five steps. (1) The challenge is first put down in writing. (2) Employees then craft a supposedly high‑quality list of questions meant to elicit adequate answers to that challenge. (3) The relevant vendors are identified and the compiled RFP document is dispatched to them. (4) Each vendor dutifully provides truthful, exhaustive answers, which are gathered and collated. (5) Finally, management performs an objective, accountable evaluation of the responses in order to pick the best vendor.
Unfortunately, every step of this mythos is wrong—or delusional.
(1) Writing, with clarity, the precise problem the software should address is harder than writing the software itself. For the uninitiated, this may sound puzzling. Yet most successful software ventures arise not from superior solutions to known problems, but from identifying a new problem altogether. Framing the problem is an open-ended task. Formal processes rigidify thinking and reduce the quality of the findings. Out of the near-infinite ways to look at a situation, one must pick the angle that lends itself to implementation. Hundreds—possibly thousands—of factors must then be technically integrated.
Corporate incentives further undermine the effort. The growing complacency of an aging executive may well be the crux of the company’s problem, but no one will write this in a semi-public memo to be shared with third parties. Such a memo will carefully steer clear of anything that could be construed as an insult to anyone’s professional capabilities within the company. Inevitably, the memo beats around the bush—protecting egos rather than the company. Far from mere etiquette, this guarantees the selected vendor will chase a solution that has little to do with the actual problem.
(2) Having employees enumerate RFP questions invariably produces material worthy of sketch comedy. Consultants add extra hilarity points. While the goal should be questions that elicit discriminative answers, the undertaking devolves into something else entirely. Questions are produced to impress peers. Those who lack smart questions compensate with volume. Questions from past RFPs are recycled ad infinitum.
Inane questions are bad enough; hundreds of them can derail a nascent initiative. Worse, they warp vendor offerings. Most such questions are requirements in disguise. Far from open-minded inquiry, those questions frame and severely restrict what counts as a “solution”. Those questions usually read, “Can the vendor guarantee that end users will be able to do X while interacting with Y?” Answering no guarantees immediate disqualification, even when “doing X while interacting with Y” is provably a terrible idea.
More generally, the there are no stupid questions mindset belongs in primary schools. Every question steers the initiative. Inane questions yield inane answers. Moreover, every question consumes a fraction of the mental bandwidth that the management can dedicate to the initiative. As executives’ attention is spread thin, it takes only a few bad questions to have them pick a vendor on irrelevant trivia.
(3) Shortlisting vendors lets someone—often not even an employee (e.g., a consultant)—make the most momentous decision of the RFP with zero accountability. In practice, the person who shortlists the vendors implicitly picks the winner. The rest is merely decorum, dictated by corporate etiquette.
While vendors may be numerous, shortlisting feels necessary only because the question list is unreasonably long. Indeed, processing that many answers for more than a handful of vendors is a daunting task. That burden is a runaway consequence of a dysfunctional approach.
(4) Vendors answering RFPs have no incentive to be truthful. The vast majority of questions beg for a specific answer: Can your solution do X? Any vendor worth his salt knows exactly what answers are expected and responds accordingly. Expecting the vendor to stick to “the truth” is delusional. Software is complex; enterprise software even more so. Nearly all questions admit a near-infinite number of interpretations. The vendor need not lie; a little imagination suffices to invent an interpretation that makes any answer “truthful”.
Even making answers legally binding will not improve their truthfulness. Enterprise software is obscure enough that almost any statement can be construed as truthful—especially with capable lawyers involved.
Such agreements—and other vexations imposed on software vendors—make the situation worse for the company. Good vendors, to some extent, pick their clients. Good vendors walk away when a prospect makes unreasonable requests. Bad vendors—desperate because their technology is found wanting—gladly embrace whatever bureaucratic insanity is pushed onto them; it is their only competitive edge. Thus, companies that take RFPs “seriously” practice adverse selection. Good vendors bow out; bad vendors persevere.
(5) Finally, a software vendor must be picked. By then, a huge pile—slanted answers meeting irrelevant questions—has been collected. All this assumes the original problem statement was correct—a premise of dubious merit. The RFP perspective supposes upper management can, by sheer intellect, pick the best vendor. Nonsense. In practice, it is usually worse.
Often, positive selection never happens. Vendors are vetoed out one by one—until only a single (viable) vendor remains. A veto requires only a sufficiently influential manager to declare himself “concerned”. Whether the concern is reasoned or merely a preference matters little. So long as he doesn’t veto multiple vendors at once (perceived as abusive interference), his objection grants him de facto veto power. Larger companies are risk-averse. Backing an initiative flagged as risky by a peer puts one’s career in jeopardy if it fails.
This final selection offers no guarantee of a good vendor. On the contrary, mediocrity is favored. A good vendor is likely transformative. After all, substantial returns on investment require departing from the status quo. Such departure invariably antagonizes barons whose fiefdoms are challenged. Those barons will vehemently oppose such vendors. A mediocre vendor is anything but transformative: he embraces the status quo and ruffles no feathers. Unfortunately, this guarantees that, at best, such a vendor delivers little or no return.
Thus, RFPs are deeply dysfunctional. They also generate considerable busywork: hundreds of questions to write and thousands of answers to survey. Hence consultants are hired to drive RFPs160. Invariably, consultants make things worse: the company’s excess busywork is their source of income. Inflating the busywork is sold as “process excellence”. Obviously irrelevant questions are retained to ensure “end-to-end coverage.” Shortcuts that would expedite the process are dismissed as “unsafe” or “untested”. At the end, consultants almost invariably lean toward the vendor offering the richest consulting prospects.
Adversarial market research
Adversarial market research directly addresses two fundamental issues in software projects: adverse incentives in vendor–client dynamics, and massive information asymmetry favoring vendors. It is less bureaucratic—and thus leaner—and more rational—and thus likelier to match the right vendor to the right challenge.
The method works as follows161. As the manager recognizes a software opportunity beyond the company’s resources, he asks one vendor—typically the one that triggered the realization—to produce a short problem statement on his behalf. The problem statement should focus exclusively on framing the challenge and the company’s specifics—not the vendor’s solution.
If the vendor cannot resist teasing his offering, the manager should simply edit out those parts; the document is short anyway. If specifics—such as figures characterizing its supply chain—are too confidential, make them vaguer, e.g., by rounding numbers. The goal is to preserve the essence of the challenge, not its fine print. The resulting memo neither names the originating vendor nor contains overly sensitive information.
With this short document in hand, the manager reaches out to a few of the vendor’s competitors. Guesswork is acceptable in picking those peers, provided they are plausible alternatives. For each competitor, the manager sends the memo, noting that it is a rough draft, that he may have misframed the challenge, and that he seeks expert guidance to reformulate the problem statement. He also requests two further items from the vendors—including the original. First, an informal back-of-the-envelope proposal aligned with the vendor’s revised problem statement. Second, a list of at least three relevant rivals, ranked by the vendor’s own assessment. Each listed rival must be accompanied by a few paragraphs explaining its perceived strengths and weaknesses.
As rivals’ shortlists reveal new competitors, those are added to the survey. Any vendor that fails to deliver these modest items within two weeks is removed from the survey. The survey ends when the manager circles back to the same vendors and no stone seems left unturned.
Once the survey has ended, the manager collates the results and creates a short file for every vendor. Each file includes what peers have said, together with the vendor’s problem statement and its associated back-of-the-envelope proposal.
Finally, with at most two colleagues, the manager reviews the results to pick a vendor. The selection is a matter of judgment and cannot be rigidly formalized. Naturally, peer opinion should play a major role in assessing each vendor. Furthermore, the quality of its problem statement and of its rival descriptions should be treated as telltale signs of competence—or incompetence if they are poor. Back-of-the-envelope proposals are used as tiebreakers. For auditability, the survey result are archived alongside a short memo explaining the reasoning behind the final selection. Afterward, the company resumes its regular purchasing process, which will likely entail obtaining a detailed offer and a full-fledged contract from this one vendor.
As stated previously, adversarial market research addresses the shortcomings of traditional methods such as RFPs (requests for proposals). Let us examine this claim, starting with information asymmetry.
Letting the company compose its own problem statement, as with RFPs, is a recipe for getting vendors to lie through their teeth to make their solutions appear to fit the problem statement. As enterprise software vendors know far more than their clients about their real capabilities, it is almost impossible to sort truth from pretense. It is unreasonable to expect any vendor to fundamentally change its technology to suit a client. Good software technology takes years to build162. With each client, the vendor may do a little better than in past initiatives, but the process is tediously slow and incremental.
This is why it is critical to let the vendor compose the problem statement. When the vendor holds the pen, it has no incentive to craft a problem statement that misfits its technology. On the contrary, it will shape the problem statement as close as possible to the solution it possesses. Furthermore, vendors—except perhaps the most incompetent—are vastly more knowledgeable than the client in framing the problem. Indeed, what the client sees as one problem may be a series of subproblems better addressed separately with distinct technologies. Conversely, this “one problem” may be merely the tip of the iceberg, requiring a broader perspective and an entirely different technological approach. A competent vendor will demonstrate mastery by presenting a convincing rationale for viewing the challenge through the right lenses.
Letting vendors expose their peers addresses the selection bias that plagues RFPs. The initial vendor selection is one of the most important decisions in the whole initiative. Early elimination of a few vendors all but guarantees a privileged vendor will be selected. So long as the survey includes half a dozen vendors or more, few will question the adequacy of the selection. This fragility is an open secret and is abundantly exploited in real-world RFPs.
By contrast, it is much harder to game the aggregate list produced by the vendors. Each vendor has its own biases and may intentionally omit a peer deemed too threatening. Yet the lack of coordination makes it difficult to single out any one vendor. Moreover, every vendor knows that its projected expertise is at stake. It will be assessed by the quality of the peers it provides. Returning to the client with a shortlist of pseudo-rivals—either incompetent or irrelevant—marks the vendor itself as incompetent.
Some vendors will produce all sorts of excuses for why they cannot provide a list of peers, let alone qualify them. Yet in a free market, a company cannot plausibly be “state of the art” while wholly ignoring what its rivals are doing. This is doubly true in software, where progress is extremely incremental. Vendors oblivious to their peers are common; it is a telltale sign of utter incompetence and should be treated accordingly.
Bypassing that entire question-and-answer game is a feature, not a bug. Those questionnaires invariably devolve into a bureaucratic monster. Irrelevant questions and requirements in disguise end up dominating the process. They frustrate the good vendors and empower the bad ones. There is simply no value in playing this game. Incentives are wrong, and neither the inputs (client’s questions) nor the outputs (vendor’s answers) can be trusted.
Similarly, the insistence on back-of-the-envelope estimates is intentional. Enterprise software is not an exact science. A competent vendor does not need a lengthy audit to draft a reasonable estimate of the total cost of ownership for a solution of its own making. Inability to provide such an estimate should be seen as a telltale sign of incompetence. At a minimum, the estimate should cover the vendor’s fees, fees to supporting parties, and the cost of internal resources dedicated to the solution. One of the most frequent shenanigans in mis-estimating costs is turning a blind eye to ancillary costs beyond the vendor’s fee. While this process cannot prevent a vendor from omitting a few lines, the lack of cooperation between vendors ensures fairly extensive coverage of expected expenditures.
By contrast, traditional RFPs coerce vendors into mind-boggling lists of inane requirements. Good vendors know that weeding this nonsense will take considerable effort and that, more often than not, they will end up implementing numerous capabilities of dubious merit. As a result, vendors invariably inflate their fees when confronted with a traditional RFP. They will also gladly turn a blind eye to every cost not directly attributable to them unless it is explicitly mentioned in the RFP.
Finally, peer assessment is the one piece of information that truly separates wheat from chaff. Every piece of software has strengths and weaknesses, but no vendor—except perhaps the most incompetent—will willingly expose the worst flaws of its own solution. When pressed, minor concerns are put forward as a show of fake humility. Misdirection is straightforward when dealing with something as complex and opaque as enterprise software. Rivals, by contrast, will not hesitate to point out the bad and the ugly. Given the nature of enterprise software, accusations or disparaging statements are unnecessary. Instead, a vendor can simply nudge the client to examine a few well-chosen aspects of a rival’s offering. Moreover, bringing relevant items to the client’s attention is, in itself, a demonstration of competence.
Adversarial market research empowers good vendors. They can train the client in the art of assessing vendors. They can frame the challenge to maximize returns when deploying the solution. They can offer lower prices, as they do not have to wade through overbearing bureaucratic insanity. The only downside is that it often demands more fortitude than complacent managers possess, as they are unwilling to ruffle the feathers of their own bureaucratic apparatus.
Once a vendor passes this adversarial survey and is selected, work returns to deployment proper: setting up raw extracts, authoring the first numerical recipe, and dual-running until no insane lines remain.
Scoping the deployment
In practice, “the same scarce resources” means a shared inventory pool; shared upstream capacity (molds, MOQs, rebate ladders, credit lines); a shared transport bottleneck (lane, port pair, vehicle class); and—routinely forgotten—a shared pocket of managerial attention. When embarking on a new supply chain initiative, the first task is to clarify its ultimate purpose and the boundaries within which it will operate. A company might, for instance, be interested in improving production scheduling or reorder quantities—or it might wish to unify several decision processes that have historically run on fragmented spreadsheets. Yet in all cases, the transformation must be made explicit at the outset: which decisions will be altered, who will own them, and how these decisions will be carried into daily operations?
One critical insight: an initiative aimed only at improving “visibility” or producing incidental metrics cannot, by itself, capture lasting returns. All those visibility efforts—even the most sophisticated dashboards—remain hollow if people in the organization carry on with the same step‐by‐step, manual decisions that predated the initiative. Conversely, once the goal is stated as a system of intelligence—one that regularly outputs and enacts tangible actions like purchase orders or capacity allocations—every requirement becomes simpler to articulate. This focus on producing—and trusting—a daily or weekly decision in a live environment forces the company to align roles and resources around a deliverable anchored from the start in the economic realities of the business.
A workable rule of thumb follows: If multiple stores draw from one distribution center, include all those stores in one go; otherwise the engine will arbitrage across spreadsheets rather than across options. If two factories or two DCs (Distribution Centers) are truly disjoint—no common inventory pool upstream, no shared trunk lanes downstream, no shared buying team—stage them in separate waves. If a supplier’s mold, MOQ, or rebate ladder binds items together, bring the whole family into scope at once. If a trunk lane or port pair is the bottleneck, scope every flow that rides it. If the same planners must arbitrate the same budget or credit line, keep their domain together: managerial attention is as finite as pallets and hours.
Deciding how broad or narrow the scope should be is no less important. An initiative might succeed in automating replenishment for a single plant but inadvertently shift problems elsewhere—such as creating bottlenecks downstream or mismatched pricing upstream. When the objective is sound daily decisions, looking at a single product line or a handful of warehouses often proves misleading. While it may appear “safer” to start small, deep structural interdependencies frequently render these localized pilot runs non‐representative. A narrowly defined project can be so constrained that the company never sees how the initiative performs under the true complexity of its operations. On the other hand, an overly expansive scope can sink the project under years of data‐integration delays. The right balance, in practice, captures all significant flows that compete for the same resources—inventory, transport, or production facilities—while avoiding optional detours into functions with little impact on supply chain decisions.
Equally crucial is establishing a minimal set of well-defined roles. Despite the myriad job titles found in a large supply chain, only a handful of positions need to be responsible for supporting an automated decision process. The Data Officer maintains the steady extraction of raw transactional data—ideally an unfiltered dump—so subsequent preparation can remain flexible and transparent. A Scientist then takes ownership of the technical and analytical stack leading from raw data to the final recommended decision. A day-to-day Flow Manager—formerly a planner in pre-automation routines—contributes contextual knowledge, challenging outputs that seem suspect and clarifying on-the-ground constraints. Finally, a supply chain executive stands ready to arbitrate when the initiative meets roadblocks—ensuring that if a new distribution center or software tool is introduced, the modeling behind the decisions is revised accordingly rather than left to drift.
All these elements—explicit decision focus, well-chosen scope, properly defined roles, and a finite but firm onboarding schedule—pave the way for a smooth initiative. They also make it evident that the object of interest is not a formula or a theoretical optimum, but a living recipe, updated and rerun in a shifting environment of suppliers, customers, and operational changes. Once that foundation is set, the next step is to examine its core: the numerical recipe at the heart of the system of intelligence, translating every relevant piece of data into the concrete, everyday decisions that steer the flow of goods.
This section revisits how to frame a supply chain initiative to produce the daily (or weekly) decisions needed to run a supply chain while it secures the “soft” elements (roles and leadership) that ensure alignment with corporate goals.
The numerical recipe
For deployment—and only for scope—the numerical recipe is best seen as a contract at the boundary of the initiative. It states in code three items scoping must settle upfront: the decisions the engine may emit (and the ledgers they write), the options it is allowed to consider (and those kept exogenous), and the shared resources it prices internally versus reads as external shadow prices. Everything else—forecasting stance, decision theory, debugging “insane” lines—has been treated in earlier chapters and is not repeated here.
First, the decisions: a scoped recipe names its emissions precisely. For replenishment: lines inserted into the purchase-order or transfer-order tables; for production: start times and quantities on a finite set of work orders; for pricing: posted price changes on an agreed cadence. Emissions must be idempotent and stateless: the same inputs produce the same outputs. The write-back can be reconciled unambiguously by a decision identifier and a timestamp. This narrow declaration anchors scope in reality: if a decision cannot be written back to the system of record, it is out of scope.
Second, the options: scope fixes what the engine may vary and what it must take as given. If the initiative covers DC-to-store dispatch while pricing remains out of scope, the price is read, not chosen; its effects enter the objective via the firm’s money ledger. If overseas ordering is in scope but carrier selection is not, transport modes arrive as costs and lead-time distributions, not as choices. Stating admissible moves prevents a relapse into spreadsheet improvisations that expand scope informally and erase accountability.
Third, the shared resources: scope is sound only if every binding resource inside the perimeter is either decided here or priced here. Inventory pools, mold families under an MOQ ladder, dock throughput, port-pair capacity, buyer credit lines, and even managerial attention are typical binders. When a binder straddles the boundary, the perimeter must ingest and honor a shadow price supplied by the neighboring domain. Absent that price, the engine arbitrages across spreadsheets rather than across options. Conversely, once a binder is brought inside, the external shadow price is retired; the coupling is internalized.
A scoped recipe also carries attribution. Its window of responsibility—from the instant a commitment immobilizes a resource until the next–next opportunity to reassign or liquidate it—is explicit, with terminal valuations and ratchets stated. Within that window, the engine bears coin-denominated consequences (lost-sales penalties, overage costs, congestion, markdowns) for its own emissions; beyond it, costs are priced as terminal values rather than chased indefinitely. This boundary prevents double counting across initiatives and the familiar blame game between neighboring teams.
Scoping requires halting and reversibility at the perimeter. The engine must stop itself when inputs are stale, when exogenous shadow prices go missing, or when dossier-level sanity checks fail. Promotion to unattended mode depends on these halts being rare and diagnosable. Reversibility is achieved through statelessness and deterministic replays: yesterday’s run can be reconstructed from versioned code and versioned extracts without pleading “data drift” or “user overrides”.
The data diet follows from scope. Tables are mirrored, in full and schema-faithful, on a daily cadence (or faster if the decision cadence demands it). No “helpful” upstream massaging is permitted: filters, imputations, and joins that matter to the decision live inside the recipe, where they can be audited alongside the economics. Minimal whiteboxing—columns that expose the main drivers of each emission in coins and constraints—also belongs to the contract. This dossier lets finance and operations see at a glance why a line is sane.
Two short examples make the boundary concrete. A DC-to-store replenishment recipe scoped to a single DC emits transfer orders for all stores fed by that DC, varies quantities and release times, reads prices and planograms as exogenous, and internalizes the DC’s labor and dock capacity. If another DC draws from the same upstream pool, the recipe must either ingest a shadow price for that pool or expand the perimeter to include the sibling DC; anything in between invites cross-DC arbitrage through spreadsheets. An overseas buying recipe scoped to a supplier family emits purchase orders with lot construction, internalizes MOQ/rebate ladders and the shared mold, reads carrier choices as priced inputs, and attributes downstream costs through a terminal value at handoff to the import DC. Pricing or store allocation remains out of scope until a subsequent wave.
Finally, scope is expected to evolve. The contract provides the seam: as neighboring domains are brought inside, exogenous shadow prices are replaced by endogenous decisions without rewriting the core. In deployment practice, this is how scope “clicks together”: the border is not a rhetorical slide but executable code that names emissions, admissible moves, binders, prices, and halts. The next subsection examines how these borders interact with the rest of the system.
The data pipeline
Scope becomes real only where it meets the applicative landscape. Modern firms run on silos—ERP for transactions, WMS for locations, TMS for carriage, CRM for orders—each with its own schema, cadence, and folklore. Scoping an initiative is therefore not an abstract drawing of boundaries; it is the decision to cut a clean, durable seam through those silos and to uphold it with a data pipeline that is thin, honest, and reproducible.
The numerical recipe defined in the previous subsection supplies the logical perimeter (emissions, admissible options, shared binders). The pipeline is its informational counterpart. It should mirror the tables required by that perimeter on the cadence implied by the decisions. In practice, a daily closing bell—an explicit cutoff (in UTC) after which extracts are taken—beats vague “near-real-time” ambitions that blur provenance. Freshness is an economic choice: accelerate only when the expected uplift in coins justifies the added friction; otherwise, favor a clean, once-a-day snapshot that everyone can replay identically. Partial extracts and “pilot slices” are counterproductive: they mask scale effects, conceal semantic quirks, and create a false sense of correctness on toy subsets. The perimeter you intend to optimize is the perimeter you mirror.
Each extract must be immutable and self-describing. Package tables with a small manifest that records the as-of watermark, per-table row counts, min/max business timestamps, and a schema fingerprint. Stores are append-only: yesterday’s files are never edited; retroactive corrections appear as new facts in today’s dump. This discipline—together with time-travel—guarantees bit-for-bit reproducibility: a run can be replayed months later without pleading “the data moved”.
Deletions and late corrections demand explicit treatment. Prefer full snapshots over clever deltas; when that is infeasible, normalize any upstream change-data-capture into an append-only ledger inside the perimeter (tombstones for hard deletes; versioned rows for updates). The goal is uniformity: the recipe reasons over a single, growing history of events, not over a moving target of in-place edits.
Access follows least privilege: a read-only service account pulls extracts; the initiative never writes back into systems of record through the pipeline. Minimize personally identifiable data; most supply-chain decisions carry no legitimate need for it. Keep the seam thin: extracts are facts, not opinions; transformations that change decisions belong inside the recipe, where they are versioned with the economics.
Historical depth should span at least one full replenishment cycle and, whenever seasonality matters, two seasonal peaks. More is better, but “enough to learn” suffices: an honest three to six quarters is enough to let the recipe converge; backfills can extend depth later without derailing deployment. External feeds—public holidays, price indices, weather for short-haul perishables—enter as first-class tables with their own cadence and provenance. Exploit internal records first; externalities pay only when their signal is demonstrably stronger than the noise they introduce.
“Helpful” upstream massaging—bespoke filters, silent imputations, pre-joined views—turns the seam into fog. Transformations that matter to decisions must live inside the recipe, where they are versioned alongside economics and can be audited with the resulting emissions. Records remain hypotheses about reality; keeping them raw preserves provenance and lets defects be traced back to their source system rather than to an undocumented cleansing script.
Because applicative silos will not be re‑platformed to suit the initiative, scoping must respect their couplings without attempting to dissolve them in advance. When a binding resource sits inside the perimeter, its constraints are internalized; when it straddles the boundary, its influence arrives as a shadow price from the neighboring domain. Attempting to “simplify” by deleting inconvenient data—or by isolating a sliver that no longer competes for shared inventory, capacity, lanes, or attention—reintroduces the very arbitrage across spreadsheets the initiative is meant to retire.
Technologically, the pipeline should be boring: append-only snapshots, immutable files, manifest-guarded extracts, and time-travel to replay any past run bit-for-bit. This modest discipline makes halting and reversibility—required by the recipe’s contract—effective in practice. Storage is cheap; premature summarization is not. The only worthwhile complexity sits where coins change hands: inside the recipe, not in the plumbing that feeds it.
Once this seam is cut—faithful mirroring on one side, auditable transformations inside the recipe on the other—scope can expand without upheaval. New domains are brought inside by replacing external shadow prices with endogenous decisions; nothing else in the pipeline needs to be reinvented. The next subsection assigns the few roles needed to keep this seam alive day after day.
Roles in the initiative
Once scope is clarified and focused on a daily production of decisions, the next critical step is to define responsibility for each part of the process. Although companies often maintain many specialized roles across supply chain, only a few are necessary to support an automated decision-making workflow. Too many roles simply dilute the ownership of the final outcome. By contrast, a small, well-defined set of roles ensures accountability and keeps the flow from raw data to final decisions clear and traceable.
The Data Officer sets up and maintains the data pipeline. This role sits at the intersection of IT and supply chain but remains firmly rooted in IT’s discipline. The Data Officer’s main objective is to deliver raw, unfiltered extractions from the systems of record, such as the ERP, WMS, CRM, or other legacy applications. Delivering these datasets daily or weekly in reliable form is harder than it looks: minor schema changes, partial upgrades, or usage quirks routinely break extraction jobs. Once stabilized, the pipeline requires minimal ongoing effort. Critically, this role should refrain from attempting to “fix” or “interpret” the data. Whenever a serious anomaly appears—like negative on-hand inventory or missing transactions—the Officer should notify the relevant teams but not overwrite or massage the underlying records. Thus the initiative confronts the real data, not an unknown transformation.
The Supply Chain Scientist owns the entire recipe from raw data to daily decisions. This role is the logical core of the initiative. While the Data Officer concentrates on extracting data, the Scientist focuses on translating every cost driver, operational constraint, or strategic preference into the recipe itself. Unlike a typical data analyst or “data scientist”, the Supply Chain Scientist is personally committed to making the process production‐grade. If a recommendation is blatantly incorrect, it is the Scientist’s responsibility to track down the source of the error—be it a mistaken modeling assumption, an overlooked lead‐time constraint, or a mismatch with the company’s current financial objectives. Over time, the Scientist codes new insights directly into the recipe, keeping it aligned with shifting operational realities. Although the Scientist uses statistical tools, optimization engines, or simulation techniques, his guiding principle is to deliver workable daily decisions. The difference in mindset is crucial: dashboards or performance metrics may drift out of relevance; daily, automated decisions cannot.
Next comes the Executive Sponsor, who provides top‐level guidance and arbitration. Supply chain decisions inevitably reflect the company’s broader economic calculus, and the Scientist will sometimes need to make modeling assumptions—about costs, margins, trade‐offs, or vendor terms—that only senior management can validate. A single warehouse might request premium freight for faster restock, for instance, but doing so may conflict with the goal of reducing air shipments company‐wide. The Sponsor clarifies how trade‐offs should be prioritized if there is disagreement. He also ensures the project has the necessary political and financial support. If the Scientist uncovers a new factor (like leftover product warranties or legal constraints on cross‐border shipments) that calls for a deeper process change, the Sponsor’s direct involvement helps clear organizational hurdles promptly.
Lastly, a Flow Manager (formerly a planner or inventory manager in older processes) remains essential. Although the daily decisions eventually become automated, real-world complexities never vanish. The Flow Manager provides the day-to-day operational insights that data tables cannot capture—for example, which supplier tends to deliver partial shipments, or which packaging lines show subtle quality issues during peak months. In the first months of dual-running the old spreadsheet and the new recipe side by side, the Flow Manager spots decisions that look “insane” and relays these observations. Over time, as automation proves stable, the Flow Manager’s job transforms. Instead of micro-checking each stock level, he focuses on broader relationships with customers and suppliers, handles unexpected disruptions such as a flooded warehouse or an overnight regulatory shift, and channels that knowledge back to the Scientist. Free from endless spreadsheet updates, he acts as a network-wide steward of the flow, heightening resilience and responsiveness across supply chain.
In brief, ownership reduces to four verbs: the Data Officer sources facts; the Supply Chain Scientist emits the daily commitments—and halts the engine when confidence dips; the Executive Sponsor arbitrates trade-offs and boundary disputes; the Flow Manager monitors field flow and escalates anomalies.
These four roles fit within a surprisingly compact structure, even in large corporations. One person may serve as Data Officer across multiple projects, and a single Scientist can manage hundreds of thousands of SKUs. Clarifying at the outset that each responsibility maps to a single, readily identifiable domain reduces cross-department friction and avoids the endless hand-offs and partial accountabilities that often paralyze more ambitious (and less focused) projects.
Once these roles are assigned, the path is set for a tangible, roughly six-month onboarding cycle—from data extraction to a stable, production-grade process.
Onboarding
Onboarding is the passage from scoped intent to unattended emissions. Once the perimeter, the roles, and the seam of daily extracts are in place, what remains is execution with explicit gates: give the numerical recipe a daily rendezvous with reality; proceed only when readiness criteria are met; and retire manual routines without interrupting the flow.
This section does not revisit scope, roles, or economics; those have been settled above. Its concern is cadence and control: what must become true, in what order, for the system of intelligence to take charge safely. Dates follow from criteria—the calendar is a consequence, not a plan.
The pages that follow take this path: a brief timeline naming the gates; the daily practice that gets us there; the mechanics of progressive rollout and reversibility; and the stance for long-term upkeep. With these pieces, the transition from spreadsheets to unattended decisions becomes routine rather than heroic.
Timeline overview
Once scope and roles are in place, onboarding proceeds through three stages that, in most firms, fit within roughly six months. This horizon is an empirical expectation, not a promise: the calendar reflects organizational cadence more than compute, and a slip of a few weeks typically signals governance or scope issues rather than “hard” technical limits.
The first stage ends only when daily data extraction is complete and reliable. The perimeter of tables required by the numerical recipe is mirrored schema-faithfully at a clear daily cutoff; extracts are immutable and self-describing; past runs can be replayed bit-for-bit; and upstream quirks are handled without “helpful” massaging that would blur provenance. Halting heuristics are wired so that the engine stops itself when inputs are stale or missing. In practice, this stage is complete once the seam has run cleanly and reproducibly for a sustained period, with defects traced and fixed at the seam rather than patched downstream.
The second stage ends when the recipe ceases to produce insane decisions. With the pipeline stable, the numerical recipe runs daily next to the incumbent routine. Each emission is accompanied by a dossier that whiteboxes its main economic drivers in coins. Day after day, surprises are fed back into semantics, constraints, or prices until glaring missteps vanish. The criterion here is not perfection but the absence of insanity: the dual-run must sustain zero insane lines, and any remaining differences from the manual routine are matters of marginal economics, not logic errors.
The third stage ends when trust suffices for production use. The very same recipe now writes back within a bounded perimeter, guarded by halting heuristics and reversible by design: emissions are stateless and idempotent; rollback and replay are routine; and documentation enables another Scientist to diagnose and revert within hours. Instrumentation is convincing enough for finance and operations to audit decisions without meetings: dossiers are legible, attributions are clear, and the window of responsibility is explicit. After a short, clean unattended rollout over part of the perimeter, promotion ceases to be a leap of faith and becomes a formality.
These gates exist to prevent pilot purgatory. If a stage lingers, escalate the cause—unfinished extracts, scope creep, or unowned prices—instead of stretching the calendar. Conversely, rushing past a gate without meeting its conditions merely displaces risk into production. The pages that follow unpack the day-to-day work inside each stage without rehearsing scope or roles already settled.
Iterative deployment
Deployment operationalizes the experimental optimization method introduced in Chapter Engineering: expose the numerical recipe to the real seam every day; let falsification drive the edits; and progress only once the resulting decisions are sane. The aim here is practical: how this stance is conducted in production, not a reprise of its theory (cf. Chapter Engineering, Experimental optimization, and Whiteboxing).
Daily dual-run
Each day at the closing bell, run the recipe end-to-end on the full perimeter, with the same cutoff and the same admissible moves it will have in production. Emit decisions but do not write them; archive both emissions and their dossiers; compare them to the incumbent routine. The run is stateless and reproducible: given the extract and the code version, the outputs are identical. Halting heuristics remain active; any halt is treated as a seam defect (missing tables, stale shadow prices, semantic drift) to be removed before the next run. Dual‑run is not a demo; it is the daily laboratory where the recipe meets reality until it stops producing surprises.
Zero insane decisions
“Insane” names any emission that violates a hard physical or legal constraint; contradicts an elementary business fact; or carries a manifestly negative expected return under the stated prices. The absence of insanity—sustained over the whole perimeter—is the gate out of dual-run and matches the second stage in the timeline above. Marginal disagreements with the manual routine do not block promotion; they are adjudicated by the ledger and settled in the next section’s rollout.
Evolving instrumentation
To make falsification fast, every emission carries a short dossier that exposes its main economic drivers—in coins—together with the few constraints that bind it. The dossier schema belongs to the recipe and evolves with it; it is versioned, replayable, and written by the same code that produces the lines. In dual-run, dossiers let practitioners pinpoint the cause of a surprise in minutes; in production, the same dossiers let finance and operations audit decisions without meetings. The mechanics of whiteboxing were developed earlier; here we keep only what accelerates deployment.
When dual-run has held for a while with no insane lines and halting is rare and explained, the same recipe can begin to write back within a bounded perimeter. The mechanics of that cutover—scope selection, safety valves, and reversibility—are addressed next under Progressive rollout.
Progressive rollout
Moving from daily parallel checks to active reliance on the system of intelligence is an exercise in risk management. The transformation must be decisive enough that the company at large no longer falls back on older manual methods at the first sign of trouble, yet cautious enough to avoid large-scale disruptions. Many organizations hesitate at this juncture—maintaining the new system as a perpetual pilot—but an orderly, firm transition keeps the initiative out of limbo.
The first scope
A common way to begin is to select a tightly knit segment of the supply chain and let automated decisions operate for that subset alone. For instance, a single distribution center might follow the daily recommendations for a specific category, or a manufacturing facility might run the automated production schedule for a single product line. Ideally, these products or facilities already have stable data and well-understood constraints. By adopting the recipe’s outputs in live operations (e.g., sending the suggested purchase orders directly to suppliers), the organization can verify that no “insane lines” appear under genuine operating conditions. This initial transition typically lasts a week or two, long enough to confirm that basic cost, inventory, and lead-time assumptions translate accurately into daily actions.
Once performance stabilizes, a broader portion of the network can be brought in. For instance, if the first wave covered one distribution center, the next may encompass all distribution centers of the same region. With every extension, the Scientist and the Data Officer must ensure that the same data-pipeline architecture scales to the added locations or product lines without new complexities. In parallel, the Flow Manager confirms that no unique local rule (such as a vendor’s strict full-pallet requirement) has been overlooked. This pattern—select, confirm stability, then expand—repeats until all relevant corners of the original scope are covered.
Manual fallback phase-out
The harder task is deciding how—and when—to retire the old manual steps. Many companies have attempted “semi-automation”, in which planners review every system-generated recommendation and adjust parameters by hand at the first hint of a shortfall. Prudent in appearance, indefinite double-checks quietly erode the gains. Nominal automation devolves into a glorified suggestion engine, ignored the moment numbers contradict intuition. Every hour spent second-guessing or retyping recommended orders is an hour not invested in tasks a machine cannot handle—such as forging supplier relationships or analyzing fresh market signals.
Truly capitalizing on automation requires retiring the older process from active use. This decision may seem to require a leap of faith. But by then, the daily comparisons should have shown that the new method rarely, if ever, produces glaring errors. When anomalies occur—supplier lead times suddenly spiking, an unannounced change to a packaging standard—correcting them in the recipe proves more systematic and faster than the spreadsheet patchwork. Because the logic is centralized in code, a single fix corrects potential errors across SKUs and locations. This clarity ultimately convinces planners and middle managers that adoption saves time without ceding control.
Handling exceptions
A dynamic supply chain inevitably encounters cases the recipe did not fully anticipate. Natural disasters might flood a warehouse, or a supplier might default. The same automation that works under normal conditions now needs a quick adjustment—either a short-term override or a modeled constraint. Immediately after the old manual process is switched off, everything that breaks feels urgent. Surges in recommended purchase orders or an inability to meet large batch sizes have to be tackled without the comfort of the old spreadsheets.
The Scientist bears primary responsibility here. Because the recipe is stateless—recomputing decisions fresh each day from the entire dataset—changes can be introduced as soon as the assumptions are clear. If a water-damaged facility must run at half capacity, the Scientist can encode that limit the next morning. The key is to integrate those exceptions into the unified code, rather than let them reappear as local overrides in scattered spreadsheets. Even after the crisis passes, the recipe retains that constraint, ready if it recurs.
Measuring outcomes
As mechanical decisions roll out across the network, instrumentation expands in scope. Previously, instrumentation helped catch erratic or incorrect lines; now it also assesses whether the predicted gains in service, cost, or reactivity are materializing. The Scientist—typically with finance and operations leads—compares actual demand to forecasts and tallies realized margins and lead times against modeled counterparts. That 360-degree view shows how well the daily decisions hold up to real-world variability. Crucially, these performance artifacts—such as accuracy measures or cost comparisons—are intended primarily for the Scientist’s iterative improvements, rather than fueling new layers of oversight. A new gap between forecast and demand might signal a subtle seasonality shift or a change in competitor pricing—an insight best folded back into the recipe’s code. If those artifacts are shared widely across departments, the organization risks returning to forecast-tweaking by committee, diluting accountability.
The mechanical rollout rests on two concurrent forms of measurement. First, the zero-insanity threshold: does the recommended order or schedule remain obviously correct for the day? Second, the longer-term actual-versus-projected checks: does the approach reflect the business’s cost realities and revenue objectives? Keeping both forms of instrumentation in the code base—visible and revisable by the Scientist—lets the Scientist locate faulty assumptions swiftly. The company retains a stable end-to-end perspective, free of the confusion arising from partial trackers or conflicting spreadsheets in separate departments.
Solidifying autonomy
Over several months of phased rollout and refined instrumentation, the initiative coalesces into a daily routine. Staff who once scrambled through spreadsheets redirect attention to more nuanced tasks or to immediate exceptions that inevitably arise in fluid markets. Confidence in the automated system grows steadily. By the time the entire scope runs under the new recipe, the manual approach is not missed. Employees see that re-litigating every reorder decision is unnecessary, while managers note that changes in demand, supply constraints, or even cost structures can be adjusted at the recipe level swiftly and consistently.
Successful automation ignites a cultural shift. Once staff realize that one piece of supply chain has truly become mechanical, they begin identifying new areas where a similar approach applies. Even if those expansions fall outside the original scope, they can piggyback on the pipeline, instrumentation, and iterative process that have already proven themselves. Such momentum spares the organization from the stagnation common after conventional software rollouts, where the system is “frozen” as soon as it goes live. Here the system’s daily outputs remain open to refinement, precisely because the capacity for quick iteration was baked into the initiative’s structure from the beginning.
Transitioning to full automation is not the final word. The next question is how to sustain and improve this new normal, ensuring the recipe keeps pace with shifts in corporate strategy or market conditions. While no one wishes to revert to spreadsheets once a well-tested mechanical approach is in place, ongoing vigilance is needed to prevent complacency. The following section explains how to keep the initiative relevant amid continuous evolution, turning each event—from everyday changes to black-swan disruptions—into an opportunity to strengthen the automated process.
Long-term maintenance
A deployed system of intelligence is never static; however polished its numerical recipe may appear, business conditions and operational realities keep evolving. New regulations may arise, pricing structures may change, and lead times or service constraints may shift in ways the automated recipe did not anticipate. An initiative that stops adapting soon becomes as obsolete as the system it replaced.
Maintenance, in this context, is not limited to software patches or data extraction fixes. Above all, it is about preserving the correctness and relevance of the economic logic encoded in the recipe. If the company opens a new distribution center or negotiates a new supplier contract, the costs and constraints underpinning daily decisions must be revised to reflect these developments. In practice, failing to maintain the recipe means either reverting to a litany of manual overrides that undermine automation or letting it produce decisions increasingly divorced from the company’s objectives.
Relentless evolution
One critical reason for systematic upkeep is that every link is constantly evolving—new SKUs are introduced, old SKUs retired, marketing campaigns appear, and unexpected disruptions force alternate routes. If the recipe stays locked in last year’s assumptions, it will miss fresh opportunities and fail to mitigate newly introduced risks.
This variability is not a flaw of automation but a feature of real-world markets. A robust initiative embraces change by scheduling regular reviews of the recipe’s assumptions and surfacing anomalies whenever the data pipeline detects large deviations from normal conditions. A surge in freight costs might call for deeper modeling of container capacities or a reevaluation of the favored transport mode. Similarly, adding advanced manufacturing lines can alter trade-offs: batch sizes once optimized can become wasteful if new lines offer cheaper, faster setups.
Finance and operations
Coordinating updates to the recipe demands ongoing dialogue between the Scientist and senior leaders in finance and operations. Day-to-day maintenance can address minor issues—a new vendor code or a recalibrated reorder policy—whereas significant changes, such as a new discount structure or a different approach to working capital, are inherently strategic.
For the recipe to remain aligned with the company’s current financial realities, it must reflect any shifting cost allocations or budgetary constraints. Because finance controls how capital is valued, the Scientist looks there for clarity on how aggressively or conservatively to scale inventory, how to treat intangible overhead (such as brand damage from a missed shipment), and how to weigh shipping delays against the expense of urgent overtime. An annual or semiannual joint review ensures the code continues to reflect the company’s strategic direction rather than drift into stale patterns. If upper management wants to prioritize expansion into new markets or slash air-freight usage by a given percentage, the Scientist cannot guess these priorities and must be informed early and precisely.
Stateless implementation
An enduring design principle for effective maintenance is to keep the recipe stateless. Each day or week it recomputes every recommendation from raw data, carrying no hidden state forward. This approach prevents errors from compounding over time. If an upstream glitch inflates today’s stock levels, tomorrow’s run—once the data is corrected—naturally reverts to normal decisions. The system can “snap back” as soon as the data pipeline recovers, without lingering ghosts of old calculations.
Stateless designs also make long-term maintenance easier. Whenever the Scientist updates a cost function or introduces new constraints, the change takes effect immediately on the next run, touching every SKU or site that needs adjustment. There is no need to weed out older stored states or to worry that a previously trained model might have internalized outdated signals. Recomputing each decision may look demanding, but modern infrastructures handle these volumes for supply chains of any size, especially once data extraction is under control.
This stateless pattern shows its worth when the system must handle surges in demand or abrupt policy shifts. If an entire region experiences an unusual spike in orders, the next daily run can incorporate it without being anchored to the prior day’s misjudgment. The recipe is flexible: it adapts as soon as conditions evolve, eliminating the temptation to rely on stale manual buffer rules once used to shore up a fragile process.
Black-swan events
Beyond everyday variability lies the risk of severe, fringe disruptions—warehouse fires, floods, pandemics, or sudden regulatory embargoes. Such black-swan events upend normal cost structures, capacity constraints, and distribution paths. While no daily recipe can anticipate events far outside typical parameters, the organization can rely on the iterative approach used during the initiative’s creation. When disaster strikes, the Scientist encodes the new or temporary constraints directly into the recipe: a site might run at half capacity or be bypassed entirely; an emergency logistics route might be assigned ten times the usual shipping cost. Once these short-term adjustments are made, the system runs again, giving an immediate view of how the new constraints reshape allocations or reorder lines.
A black-swan event might also require deeper changes—building alternative flows or drastically revising lead times for an entire region. The principle remains: the recipe is the single point of integration, so the newly discovered constraints or extreme costs are introduced there. This approach is more resilient than ad hoc firefighting through manual overrides scattered across spreadsheets and departments. Each fix, once validated, becomes part of the code base and ceases to rely on ephemeral human vigilance.
Crucially, an extraordinary crisis can become an opportunity to deepen the company’s risk awareness. If a supplier default reveals that 90% of a key component was single-sourced, the recipe can be permanently updated with rules that prevent overexposure to a single vendor. The same instrumentation used for daily checks can surface such vulnerabilities. In time, lessons from turmoil elevate the baseline, ensuring the system emerges more robust.
Long-term maintenance, therefore, is not only vigilance against normal drift but also a structured response to sudden shocks. Both daily updates and urgent fixes tie back to a single stateless recipe that remains the core of an automated supply chain. Once a company becomes accustomed to this cycle—small increments, periodic recalibrations, and crisis-driven overhauls—it finds that the automated flow stays current. The initiative proves more stable and flexible than any patchwork of spreadsheets because it has a built-in mechanism for continual renewal.
At this juncture, the initiative has a stable architecture, daily operations run under automated decisions, and the company is ready for both routine changes and big surprises. What remains is to assess how this redefines roles, organizational structures, and job descriptions: a well-maintained recipe does not merely replace old processes; it transforms the supply chain workforce.
Organizational changes
Organizational shifts are unavoidable once automated decisions flow daily through the company. Though the mechanics of a well-defined scope and an iterative rollout may feel purely technical, people across the business soon realize that many tasks—once handled by planners or middle managers—now proceed with little human intervention. Employees accustomed to gathering data, reconciling conflicting indicators, and juggling last-minute overrides discover that the new process leaves fewer gaps to patch. Some welcome the relief and focus on thornier supply chain problems that resist automation. Others sense a loss of relevance and resist the change, citing alleged “exceptions” or insisting the system “cannot capture every detail”.
Yet once laid out explicitly, these details can be encoded into the recipe. If, for instance, a planner tweaked reorder quantities for a vendor that consistently shipped partial pallets, that nuance belongs in the code. Once added, the recipe no longer relies on last-minute fixes. Freed from perpetual firefighting, employees can extend the scope of the initiative or strengthen relationships with partners. Over time, the company must acknowledge that many day-to-day routines have changed. Traditional planners, accustomed to scanning spreadsheets for hundreds of SKUs each morning, find such manual checks redundant. Finance, which once saw forecasts as loose “best guesses” that demanded caution in budgeting, is now confronted with a system that commits capital daily unless told otherwise. The IT department, historically overloaded with backlog requests for reports or mini-applications, finds that a single well-engineered pipeline supplants many scattered customizations.
Overcoming these frictions requires a clear articulation of why the initiative demands not only a shift in software but also a transformation in who does what. If the automated decisions are trusted, the company no longer benefits from employing multiple teams to replicate or override them. Rather than stewarding endless spreadsheets, planning staff should enhance how each product is sourced or stored, refine cost assumptions, and capture new signals that would otherwise remain invisible to the algorithm. Similarly, finance executives need not revisit the same line-by-line details if the economic drivers in the recipe faithfully reflect the intended constraints. Instead, their role is to adjust those drivers when strategy changes or new resource limits apply.
This shift also calls for new forms of cooperation. A mechanical system is both transparent and inflexible: it cannot quietly paper over a mismatch between two departments. If a warehouse control rule contradicts a recent marketing push, the conflict appears quickly in the daily outputs. Ultimately, resolving such contradictions requires more than a polite meeting or an ad hoc fix. It requires the Scientist—or whichever role now owns the recipe—to incorporate the updated guidelines and then rerun the logic. For that to happen without chaos, the organization must unite behind a single automated process rather than let each team revive a personal spreadsheet whenever it dislikes the result. Those who cling to local workarounds will dilute the benefits of automation for everyone.
The most visible change unfolds in the day-to-day planning workforce. Previously, they rotated across SKUs or sites, adjusting inventory levels, double-checking lead times, and manually intervening when supply or demand changed abruptly. Now they can engage more strategically. Instead of verifying hundreds of reorder lines, they might confirm that new supplier contracts are coded correctly or investigate an unexpected surge that the recipe flags as an outlier. Their daily work becomes less about mechanical data entry and more about relationship-building, discovering new cost drivers, and anticipating shifts in the market. That evolution can feel unsettling at first. Some employees may assume their original role—manually handling each step—remains irreplaceable. They may try to keep partial manual processes alive. But once the system runs smoothly for a small subset and expands to a broader scope, the recipe’s consistency soon outperforms scattered human vigilance.
Although organizational changes can trigger discomfort, they pave the way for the company to become more agile and more aligned with genuine priorities. If a strategic initiative aims to lower transport costs or reduce overall lead times, it can be implemented directly in the recipe, then tested and measured on real orders—not just debated in a meeting. Planners gain opportunities to shape how constraints are modeled, rather than playing catch-up each time demand shifts. Executives see a clearer cause-and-effect between constraints they approve and the subsequent outcomes in inventory or fulfillment metrics. As unproductive routines dissolve, staff once absorbed in repetitive tasks can focus on forging stronger ties with suppliers, gleaning meaningful signals from the market, and ensuring that new product introductions or transitions align with the daily mechanical approach.
These organizational adjustments form the backbone of a supply chain that thrives on rigorous, automated decisions. In the sections that follow, we turn to the most direct impact on planners themselves, who evolve into “Flow Managers” within this new framework. We then examine how the Scientist role becomes central to sustaining and improving these decisions long after the initial deployment.
Planner vs. Flow Manager
When daily decisions become fully automated, the traditional role of the planner—poring over spreadsheets, adjusting reorder quantities, and reconciling conflicting priorities—gradually loses much of its daily purpose. In many ways, this shift is liberating. Freed from the burden of chasing small data inconsistencies or applying repetitive heuristics, the same person can step into the higher-value role of Flow Manager. Rather than micromanaging every SKU, they now survey the entire supply network, identify real-world exceptions, and cement collaborative ties among suppliers, carriers, and internal stakeholders.
Repetitive adjustments
Planners were typically tasked with aligning moving pieces—forecast data, supplier lead times, and local constraints—by manually reconciling spreadsheets and system-generated figures. They were the “glue” holding processes together whenever the formal software or the official forecast fell short. In an automated setting, however, the logic that once filled these spreadsheets is embedded in the recipe that produces each day’s decisions. This recipe does not merely generate theoretical numbers: it proposes or enacts reorder lines, allocations, or capacity bookings, each grounded in the company’s financial costs and constraints. Whereas before a planner might spend hours each week scouring for suspicious stock levels or last-minute shipping mistakes, the process now flags outliers automatically, prompting targeted interventions rather than universal manual checks.
As a result, the focus of the role rises to a higher level. Instead of scanning hundreds or thousands of SKUs for reorder miscalculations, the ex-planner can invest his energy in more strategic or more relational tasks. He can track subtle signals that numeric models might miss, such as a supplier’s newly acquired certification or early warnings that a carrier is struggling in certain regions. He can also assess areas where the recipe might benefit from deeper insights—such as incorporating real-time capacity data for a critical production line or capturing new discount schedules from a primary vendor. By supplying these discoveries back to the Scientist, the Flow Manager ensures that the automated recipe evolves without daily guesswork.
The state of consensus
This transformation often does away with the repetitive alignment sessions—commonly known as Sales and Operations Planning (S&OP)—in which teams formerly debated how to adjust forecasts or handle uncertain capacity. In many companies, these sessions devolve into monthly or quarterly rounds of nudging numbers, producing a single consensus figure still disconnected from operational realities. Such a process, at best, repeats known constraints and departmental viewpoints; at worst, it masks conflicts and leads to feeble compromises.
With a numerical recipe driving daily allocations or production schedules, there is no protracted debate about which forecast is “right”. All relevant constraints—lead times, cost structures, stockout penalties—are either in the system or not. When a department raises a concern (for instance, a marketing plan that will push demand spikes on short notice), the response is concrete: the recipe must be amended to reflect the true economic stakes. One might encode higher penalty costs for stockouts of certain promotional items, or factor in premium freight for sudden surges. Rather than negotiating a middle-ground forecast in weekly meetings, the Flow Manager and the Scientist ensure that operational decisions align automatically with updated models of costs and constraints. This renders the old forecasting debates largely moot—every “final figure” that shapes daily orders is anchored to the company’s economic logic, not a compromise hammered out in a conference room.
Real exceptions
Because the numerical recipe self-corrects in response to new data, daily overrides for individual SKUs become unnecessary. The Flow Manager’s vigilance instead focuses on genuinely abnormal events—cases so far beyond the recipe’s rules that they would produce decisions at odds with reality. For example, if a key supplier closes a plant for two weeks after an unforeseen technical failure, the manager identifies a major disruption. Instead of overriding hundreds of reorder lines, he elevates the matter: “Our lead time from Supplier A is effectively doubling this month, with partial shipments only.” The Scientist codes the change, the recipe reruns, and the entire system readjusts accordingly—from revised reorder quantities to different carriers, if needed. In the former manual approach, such an event would trigger days of frantic email threads and one-off spreadsheet changes in multiple regions. Now it is handled as a single change in the logic.
This shift in focus reduces friction across the organization. The Flow Manager no longer needs to defend, line by line, how he reached certain reorder amounts. Instead, he can examine the broader shape of demand spikes or supply constraints. If certain items systematically underperform service expectations, the question is not “why did you reorder too little last week?” but rather “how can we reframe the cost drivers or lead-time assumptions so that the recipe captures this item’s volatility?” True issues get resolved swiftly; old patterns of “we’ll just push up the forecast” fade away.
Strategist, liaison, and auditor
Over time, the rebranded role of Flow Manager stands at the intersection of three core missions. First, he becomes a strategist—no longer stuck in day-to-day reorder monotony, he refines the underlying financial logic, highlights future expansions (new warehouses or product lines), and proposes broadening the automated scope. Second, he serves as a liaison—communicating with sales, marketing, and procurement to gather developments the automated system has not yet captured. Finally, he acts as an auditor—spot-checking the system’s outputs for outliers, investigating anomalies, and verifying that the cost or lead-time assumptions still match real-world conditions.
None of these missions resembles the classic routine of adjusting safety stock in a spreadsheet every morning. That old pattern is replaced with continuous improvement cycles. Each cycle expands either the coverage of the automation (taking on new product families, channels, or constraints) or the depth of the model (introducing better cost approximations for shipping, new rules for batch production, or more nuanced penalty structures for delayed deliveries). The ex-planner, freed from clerical tasks, is key to unearthing which expansions matter most.
The wider team
When the Flow Manager devotes energy to strategic liaison work, the former roles of “senior planners” or “production coordinators” must likewise adapt. In organizations that cling to older structures, multiple staff may still attempt to replicate daily decisions offline. This duplication only reintroduces confusion—no two spreadsheets mirror each other perfectly. But in organizations that embrace the shift, employees who once specialized in manual reordering can become domain experts. They may concentrate on specific complexities, such as ensuring hazardous-materials shipping follows precise regulations, or forging stronger lead-time improvement deals with logistics partners.
Moreover, while S&OP or forecast-collaboration meetings might vanish in their traditional form, some communication channels remain. The difference is that these channels no longer revolve around “tweaking” forecasts. Instead, they focus on bridging the recipe’s logic with the front lines of the business—discussing new store expansions, pending product line retirements, unexpected raw material surcharges, or brand-protection strategies that, if not coded, lead to suboptimal daily outputs. Each conversation is grounded in the same integrated logic that drives every decision, and participants talk about adjusting real drivers instead of haggling over whose forecast is correct.
In sum, as automation takes hold, the planner’s role rises from daily micromanagement to that of an empowered Flow Manager. By assuming accountability for capturing operational truths—both subtle and dramatic—the ex-planner-turned-manager ensures that the mechanical decision process remains accurate, timely, and ever-improving. Freed from spreadsheet chores, his insights expand to serve the entire network, from supply partners to in-house manufacturing, forging deeper synergies than the old routines ever permitted.
The Supply Chain Scientist
The Supply Chain Scientist carries a unique responsibility: translating the organization’s evolving economic realities into a numerical recipe that steers the daily flow of goods. Once the company commits to replacing large swaths of manual planning with automated decisions, the Scientist becomes the process’s linchpin. This role stands apart from traditional business analytics or forecasting positions because it mandates both end-to-end ownership of the recipe and direct accountability for profitability. Instead of merely improving visibility or producing numerical artifacts, the Scientist aims to deliver correct, self-adjusting decisions, trusted in production day after day.
Bridging with IT
From a technical standpoint, the Scientist sits at the frontier of the company’s systems and data. Yet it would be a mistake to treat the Scientist as part of the IT function. The Scientist must be free to focus on modeling supply chain decisions, not on wrestling with a litany of data-integration chores that IT already specializes in. The Scientist does rely on a stable pipeline of raw records—unfiltered extracts of transactions, inventories, supplier terms, and the like. However, once that pipeline is set up, the Scientist’s mission pivots to coding new insights and constraints that reflect supply chain conditions. The IT group, for its part, remains the caretaker of data extractions and infrastructure.
This division of responsibilities benefits both sides. The Scientist does not worsen the IT backlog with endless demands for specialized transformations; the data officer in IT delivers neutral, daily snapshots of the source systems, letting the Scientist handle all refinements in the recipe. Meanwhile, the IT department can concentrate on its broader operational mandate—maintaining the stability of key applications and ensuring that every upstream database and data feed remains healthy. Once a problem surfaces, such as a table rename in the ERP or a failed overnight export, the two sides cooperate quickly: IT resolves the pipeline glitch, and the Scientist resumes normal operations with the updated data. There is no friction or confusion about who “owns” the transformation logic or which changes were introduced midstream.
On finance
A corollary of the Scientist’s mandate is the need to integrate financial constraints deeply into daily decisions. Cost drivers—whether freight rates, working-capital limits, or margin targets—cannot remain abstract ideas floating in slide decks. They require explicit numerical treatment so that every allocation roots the trade-offs in genuine monetary realities. The Scientist thus cooperates with finance not out of formality but out of necessity: if upcoming promotional campaigns alter expected margins, or if new capital constraints preclude particular inventory levels, the recipe must be updated accordingly.
These exchanges with finance take place more regularly than might be assumed. During the initiative’s rollout, the Scientist typically holds a series of short meetings where he presents how current cost assumptions—such as storage overhead or penalties for expedited freight—translate into daily decisions. Finance might spot hidden inaccuracies, such as an overlooked handling fee for bulk deliveries or a missed opportunity to factor discount tiers more finely. Correcting these details can shift the solution from a naive local optimum to decisions that better reflect the company’s real P&L impact. This is how the automation remains financially “sharp”—no budget line or profit center is left to guesswork.
Over time, even after the recipe matures, major operational or strategic shifts still prompt fresh rounds of modeling. A high-level directive—such as lowering the cost of capital or diminishing exposure to a class of risks—flows through finance to the Scientist, who translates it into revised penalty functions, discounts, or sourcing rules. The effect appears in the daily outputs soon after, keeping the company’s finances and day-to-day supply chain harmonized.
On leadership
Although the Scientist is neither the CEO nor the Executive Sponsor, he does hold a pivotal seat at the table whenever top leadership alters strategic priorities. As soon as the new direction crystallizes—entering a new country, switching to nearshoring, favoring premium service for certain product lines—the numerical recipe must encode those priorities. Without this prompt alignment, the mechanical decisions remain stuck in outdated logic.
A direct, brief conversation with top leadership once or twice a year can suffice. Its purpose is not to pore over micro-level code changes. Rather, it confirms that the Scientist understands the larger picture. If the company aims for faster expansions yet stricter capital usage, the Scientist might reflect this by tightening inventory budgets on slower-moving items or by introducing steeper penalties for using emergency freight. The recipe’s daily outputs thus shift where they will influence the bottom line. Because the code is stateless, no states linger to hamper the transition. Within a week of leadership’s new directive, the supply chain might already be producing purchase orders aligned with those strategic imperatives.
This pattern of swift adaptation demonstrates how strategic coherence emerges from the supply chain’s mechanical foundation: once the code’s assumptions change, every relevant decision shifts as well. Where a manual process once required months of incremental retraining for dozens of planners, a single update here instantly scales across the entire product range or site network.
Beyond two Scientists
In many midsize companies, a single Supply Chain Scientist—plus a second for redundancy—can cover entire decision streams. Yet as the initiative grows—adding more product lines or diving deeper into advanced production or distribution problems—additional Scientists may be needed. Managing a larger team requires a more rigorous approach to the recipe’s structure.
The first principle is to keep as much shared logic as possible in a unified codebase. Each Scientist might handle distinct sub-areas—one focuses on procurement constraints, another on production scheduling—yet all collaborate on a single blueprint. Just as software engineers rely on version control and code reviews, Scientists adopt similar practices to coordinate daily changes and ensure an improvement in one domain does not inadvertently break constraints in another.
Crucially, Scientists remain supply chain experts first and foremost. They are not hired to pursue abstract machine-learning innovations. If they set up code reviews, the goal is not to create a specialized IT pipeline but to keep daily decisions consistent and error-free. Each Scientist learns the rationale behind major lines of code—the “why” as well as the “what”. A master manual163 can prove invaluable, tracking the strategic motivations for each rule. When one Scientist departs or a new hire arrives, knowledge transfer is smoother and spares the group from rediscovering intricacies of cost assumptions or corner-case constraints.
By handling expansions in this manner, the Scientist’s team can scale gracefully while preserving the daily iteration pace. No matter how large the business becomes, every aspect—from new discount structures to refined lead-time predictions—can be tackled by the relevant Scientist and then merged, in an orderly way, into the master recipe.
Strict focus
Although they code extensively, Scientists are not subordinate software developers. Their success is measured by the real impact of daily decisions—lower inventory misalignments, higher margins, or faster adaptation to abnormal conditions. They do not vanish into IT’s backlog, and they are not a generic “data science” team chasing ephemeral accuracy benchmarks. Instead, they retain a sharp focus on actionable changes that reorder the flow of goods to better serve strategic and financial goals.
This grounding in the tangible separates them from departments that traffic in dashboards or once-a-quarter analyses. The Scientist must repeatedly take small, purposeful steps with code whenever the environment shifts, verifying each day’s results in production. The synergy with IT is essential but bounded at the data-extraction boundary. The synergy with finance persists, ensuring that cost drivers retain credibility even as the market moves. And the synergy with senior leadership tightens, allowing the automated process to reflect any new direction the company adopts swiftly.
By internalizing these realities, companies realize that the Scientist’s role is neither ephemeral nor disposable after the “automation project” ends. Just as the Flow Manager replaces the manual planner in day-to-day tasks, the Scientist replaces a tangled web of partial solutions with an evolving, codified approach. The necessity of strategic updates never ceases, and the Scientist stands ready to embed each shift into the mechanical logic. This continuous interplay is how a supply chain remains perpetually relevant and linked to real economic drivers.
The Scientist, in short, is a new breed of contributor to corporate operations—one whose orientation is neither purely technical nor purely managerial, but a binding of the two. Their authority is drawn not from any formal managerial hierarchy but from the everyday accuracy and adaptability of their code. In the next section, we examine in greater depth the common pitfalls that befall “data science” in corporate contexts, and why genuine Supply Chain Scientists must go beyond superficial number-crunching to attain lasting success.
The greater company
Every initiative that reshapes supply chain decisions, no matter how precisely it is scoped at the outset, inevitably presses on the corporate fabric as a whole. The numerical recipe cannot operate in isolation—its daily outputs must reflect the constraints, objectives, and cost structures of multiple departments, many of which have traditionally maintained local processes disconnected from broader concerns. While the earlier sections showed how to structure the initiative around a Supply Chain Scientist, the Data Officer, the Flow Manager, and the Executive Sponsor, the rest of the company must also learn to interact with—and ultimately trust—this new decision-making process. In the absence of a unifying framework, local, short-term pressures from sales, marketing, procurement, or other areas risk overriding or undermining the mechanical output.
Yet the real strength of the recipe lies precisely in gathering every relevant driver and expressing it daily, without hidden bias or fragmented oversight. For that to happen, each department must relinquish the temptation to maintain its own confidential spreadsheets and shadow processes. It must, instead, state its constraints and cost structures in the open so the Scientist can encode them directly. This is where upper management, and in particular the top executive or CEO, must intervene decisively. The CEO is not, and cannot be, the final arbiter of economic drivers; such details are too fine-grained for the top office. Rather, the CEO is the ultimate arbiter of which department owns each driver, and of ensuring that these owners cooperate with the Scientist rather than attempt to circumvent the system. If finance asserts that certain capital costs must be set at a specific rate, those costs become the recognized standard. If procurement stipulates that a particular strategic vendor is to be spared from cuts in order volumes, that choice is modeled explicitly. The policy must come from the relevant department, but once confirmed, it ceases to exist as an unwritten rule and is anchored in the daily decision logic.
This arrangement can feel jarring to departments that previously adjusted their numbers on a whim. When a brand manager finds that a popular product does not appear in the upcoming order plan, he may fear losing revenue or market share. In the old approach, he could scramble to override the reorder quantity in a spreadsheet, whether or not it truly benefited the company’s bottom line. In the new system, if the brand manager believes that omitting the product causes reputational damage that far outweighs any short-term profit concerns, the correct path is to escalate this factor so its financial cost or penalty is coded into the recipe. The goal is not to “blame” the mechanical process for ignoring a brand imperative; it is to embed that imperative—expressed as a penalty for stockouts or a premium for brand-building availability—in the economic drivers. From then on, the daily decisions that once led to friction will naturally align with the brand manager’s priority, so long as that priority remains truly important to the company’s overall strategy.
Similar tensions arise with marketing when a promotion is omitted from the final plan, or when a featured product proves marginally profitable and is assigned a low reorder priority. If the marketing team has evidence that a promotion will raise cross-selling of other items, it must show how the additional revenue or brand equity translates into an economic payoff. Once the added payoff is stated explicitly, the Scientist incorporates it into the code. Those marketing concerns no longer get handled through last-minute overrides or “final forecast adjustments”; they become part of the daily mechanical process—an open, documented variable that can be revised if market feedback shows the promotion did not generate the expected cross-selling. This open articulation of what were once tribal or intangible factors forces each department to decide whether a local desire is a passing hunch or a vital economic truth. It is, in effect, a game-theoretic dynamic: every department sees that partial cooperation (such as whispering an inflated demand to secure inventory) no longer yields an advantage. The final recipe checks all constraints daily; if a cost or benefit is not declared, it is not recognized. Top management’s role is to ensure each domain states its costs and gains honestly, rather than manipulating demand figures to secure resources at others’ expense.
Procurement, likewise, may initially feel threatened when the daily recipe dictates exact reorder lines and timing, effectively replacing the manual negotiations some buyers believed were their primary value. Yet the procurement function remains indispensable in revealing which vendors have reliable quality, which suppliers impose hidden conditions, and which may merit sacrifice in favor of new sourcing strategies. The procurement team does not need to overshadow the system or revert to side deals; it gains more influence by stating vendor constraints clearly. Are there minimum batch sizes, special inbound tariffs for partial shipments, or vendor loyalty discounts? Each nuance is a piece of the economic puzzle. Once coded, it shapes the recipe’s outputs, so procurement influences daily decisions through stable, rational rules rather than repeated manual rework.
The same logic applies to engineering teams and product designers who want to incorporate new components or adjust bills of materials. In the past, a design modification could blindside planning, causing last-minute shortages when the newly specified part was overlooked. Now the mechanical daily process immediately flags any unknown part, revealing that the new item lacks cost references or lead times. The design team can no longer slip changes through unnoticed; it must collaborate with the Scientist to define how the new component alters production constraints or overhead. This enforced collaboration spares the company from silent supply mismatches that once appeared “inevitable”.
Even finance—long accustomed to viewing supply chain as a cost center apart from brand or marketing priorities—is pressed to clarify how liquidity constraints and working-capital targets feed into daily reorder levels. If finance wants certain lines to run with minimal buffer stock, or signals that inventory valuations are changing, those updates translate into immediate code changes. The daily outputs no longer merely reflect a forecast and a standard margin; they express finance’s up-to-date stance on capital usage. When a top executive wonders why certain SKUs have been cut back, there is a direct line from finance’s constraints to the automated decisions—and from the Scientist’s code back to finance’s constraints. No arbitrary overrides are needed and no endless forecast meetings are held; everyone sees the cause and effect.
In such an environment, the question of blame rarely arises. If an unexpected shortfall in a promoted SKU hits the shelves, marketing will challenge the numerical recipe—not to undermine it, but to check whether the promotion’s impact was correctly included. The Scientist, in turn, checks whether marketing truly disclosed all relevant information—such as incremental brand risk or cross-selling gains. If not, the remedy is a new or revised cost assumption. Over time, these cross-department interactions evolve from friction-laden skirmishes into constructive updates to the recipe’s parameters. This model of collaboration thrives only if top management, from the CEO downward, endorses the process and stands firmly against reintroducing manual manipulations whenever a department is dissatisfied with an outcome.
Once each department sees that alignment comes from adjusting shared assumptions—not from ignoring or overriding mechanical outputs—it realizes the mechanical decisions have become the single source of truth for daily operations. That single source is not rigid or unwieldy; it adapts as soon as a cost, constraint, or margin driver proves relevant. It refuses to incorporate half-baked ephemeral guesses. This dynamic can be disconcerting in the initial months, as employees discover they can no longer quietly fix an inconvenient number in their spreadsheets. Yet after a full cycle—often a quarter or two—people recognize that fewer emergencies arise, fewer turf wars erupt, and the supply chain feels more coherent.
The initiative becomes more than a supply chain improvement; it is a mechanism for continuous cross-department synergy. Every domain within the company—marketing, procurement, engineering, finance, and operations—sees that the best way to secure its priorities in the daily flows is to articulate them openly and then to trust the mechanical process to execute them consistently. The CEO’s role is to ensure each domain truly owns the costs or benefits it claims and to endorse the Scientist’s final code as the channel through which those claims reshape decisions. Anything less breeds duplication or pushback that dilutes the new system’s gains.
Over time, a culture emerges in which no one fights the recipe’s daily numbers so long as they reflect properly stated drivers. If marketing finds brand costs understated, it raises the issue formally. If procurement sees a vendor’s reliability has improved, it updates either the penalty for stockouts or the expected lead times. Everyone gradually grasps that game-like maneuvers—such as inflating forecasts—no longer yield an advantage, because the recipe cross-checks everything with actual data. The shared impetus is to refine numerical assumptions, not sabotage them.
This shift toward a single, daily, mechanical orchestration may appear dramatic in companies long reliant on departmental autonomy. The benefits include sharp reductions in firefighting, deeper synergy across brand, marketing, and operations, and a steadier link between strategic ambitions and on-the-ground execution. The moment each department accepts that it must shape the inputs—rather than override the outputs—the initiative achieves the company-wide collaboration essential to long-term success.
Stagnation
Supply chain, as treated in this book, is an intent executed through decisions; the narrow claim of this chapter is that, as a discipline, supply chain has largely stagnated. Not because goods fail to move—they do—but because firms’ choices of what to move, when, where, and how much have not kept pace with four decades of computation. The dominant posture remains clerical: people reconcile exceptions produced by software that records facts and narrates the past but rarely places bets.
By stagnation we mean operational stasis masked by motion. Ledgers are digital and ubiquitous; reports refresh by the minute; data warehouses swell; acronyms and dashboards proliferate. Yet the core mechanics of allocation are much as they were in the mid-1980s: point forecasts treated as certainty, buffers tied to arbitrary service targets, reorder surrogates wired into weekly cadences, and a daily swirl of approvals. Computers serve as systems of records and reports, and where “automation” appears, it is usually alert theater—lists that hand the hard cases back to humans—rather than a system of intelligence that issues unattended, auditable commitments.
This stasis exacts a cost. It systematically suppresses what moves money: fat-tailed variability, couplings across time and space, and the option value of waiting. Averages stand in for distributions; local performance indicators stand in for economics; Goodhart’s law quietly turns measures into targets. The result is a clerical equilibrium that delivers continuity through buffers and overtime while letting coins leak in ordinary times but faltering under stress.
None of this is inevitable. Earlier chapters have assembled the missing instruments: probabilistic views of the future instead of point surrogates; coin-denominated valuations at firm scope; and systems of intelligence that recompute within frames, price trade-offs, and write decisions back unattended—guarded by halting heuristics, whiteboxed dossiers, and experimental optimization under dual-run. When these tools are adopted, the same assets produce better flows because the decisions change.
What follows is both synthesis and indictment. The field did not merely fail to accelerate with modern computation; it settled into a stable, mediocre equilibrium sustained by incentives that reward narratives over emissions: academic puzzles, vendor rebranding of records and reports as “planning”, consulting ceremonies, and internal “data science” that stops at insights. The organizational trap—supply chain’s second‑class status and subordination to central IT—anchors the arrangement.
This double image—a flow that keeps delivering and a discipline that stands still—sets up the paradox that follows.
The paradox of success
Supply chains are celebrated as dependable. Goods cross borders daily; store shelves are rarely bare for long; companies post solid results while expanding their networks. Judged by visible outcomes, the field appears vigorous and resilient.
Yet this outward success masks a deeper stalemate. Over decades of extraordinary gains in computing, most organizations have treated computers as ledgers and dashboards rather than decision engines. What passes for “automation” is usually a cascade of exceptions for humans to reconcile; what passes for “planning” remains a brittle blend of deterministic forecasts, reorder surrogates, and manual overrides—merely reskinned with modern interfaces. Continuity is bought less with engineered intelligence than with buffers, redundancy, and clerical vigilance.
This is the paradox: reliable delivery at scale coexists with intellectual stasis. Other domains have converted cheap computation into unattended decision making; supply chain has largely repackaged yesterday’s routines. The much-touted innovations are cosmetic—new acronyms for old artifacts, new screens for old habits—while the core remains a daily, human-mediated reconciliation of numbers.
The cost of this stasis seldom appears in a day’s snapshot because the flow does “deliver”. It becomes conspicuous during shocks—abrupt demand shifts, transport disruptions—when processes designed for compliance rather than adaptation to uncertainty falter. More damaging is the quiet, compounding bleed in ordinary times: overgrown safety stocks, underused options, and thousands of hours spent curating spreadsheets instead of improving decisions. Expensive software is acquired, but it rarely elevates decision quality; its “automation” devolves into alert theater that assigns the hard parts back to people. More insidious than any visible miss is the mounting opportunity cost: every day that variability is treated as a nuisance rather than priced and exploited is a day when surplus that modern computation could harvest simply evaporates.
Earlier chapters outlined remedies that already exist: probabilistic views of the future, economics‑anchored valuations, and systems of intelligence that emit unattended, auditable decisions daily—continuously improved through experimental optimization. The aim here is not to re-argue those instruments but to explain why a field that could adopt them largely has not. The next section traces the roots of this equilibrium—academic fashions, vendor economics, consulting theater, and corporate governance—that keep a demonstrably central function stable and mediocre.
Real-world complexity
Invoking “complexity” has become ritual apology for mediocre practice. The problem is not that supply chains are complicated—our first chapter already distinguished the essential from the accidental—but that most organizations flatten what matters economically into a handful of surrogates. Point forecasts stand in for uncertainty; a single “service level” impersonates profit; average lead times erase the tails that do the damage. The result is not a model of the flow but a tranquilized caricature of it.
Simplification is not a vice when it preserves the asymmetries that move money; it is a fault when it erases them. Variability and optionality are the raw materials of profit. When a representation suppresses fat tails, correlations, batching, or ratchets that bind tomorrow’s options, it no longer guides decisions—it conceals their stakes.
What follows is not another taxonomy of failure but a concrete vignette. A small, operational episode makes it sharp. It also bridges to the method that consistently breaks stalemates in the wild: experimental optimization (see the chapters Decisions, Engineering, and Deployment).
Operational vignette
Consider a mundane episode. A national retailer operates two regional DCs, each with a nominal receiving capacity of forty containers per day. Purchasing works on a weekly cadence keyed to supplier MOQs and rebates; the planning system computes reorder points from point forecasts and average lead times; warehouse KPIs track dock utilization and turns; stores are judged on “on-shelf availability”. No villain, only sensible local goals.
Every Monday, sixty to seventy containers arrive at DC West. The gates saturate by noon; trucks idle; demurrage mounts; a fraction of the freight rolls to Tuesday; the tail spills into Wednesday. Meanwhile, store shelves run patchy mid-week on long-tail SKUs that were “in transit” in bulk on Monday yet physically receivable only on Wednesday. No dashboard lies: service looks fine on a weekly average; the rebate ladder was captured; dock utilization peaked; inventory turns hold. Yet coins leak everywhere—idle trucks, overtime, transient stockouts—while the organization congratulates itself for “hitting targets”.
Nothing subtle is required to fix the picture; what is required is to admit the picture. In dual-run, a thin numerical recipe was introduced: lead times as distributions with fat right tails; a priced shadow for dock saturation that rises nonlinearly beyond forty containers per day; supplier MOQs as penalties rather than edicts; and, crucially, a daily assessment of orders without an obligation to order daily. The engine could emit “do nothing yet” whenever the option value of waiting exceeded the expected, risk-adjusted return.
On day one, the halt tripped on insanity—a planned Monday inbound of forty-eight containers for a forty-container dock; dossiers exposed the drivers—priced in coins—and the defect was removed: weekly batches were replaced by staggered releases that preserved rebates while flattening the inbound. Within four weeks of dual-run, demurrage and receiving overtime fell markedly; mid-week shelf gaps on the long tail diminished without raising working capital; rebate capture persisted. No heroics, no grand reorganization: the same assets, the same suppliers, the same stores—only the decisions changed.
The lesson is not that “Mondays are bad”. It is that the real complexity was economic and temporal, not algebraic. Averages hid tails; a single service metric hid heterogeneous payoffs; hard “constraints” became elastic once priced. Computers did nothing human planners could not, in principle, do—they simply did it for thousands of micro-options every day, without fatigue, surfacing the right frictions in money.
Local maximum
The episode shows why mediocre equilibria persist. Each function perched on a defensible hill: purchasing on rebates and MOQs; warehouses on utilization; stores on availability; finance on turns. The hills were real; the mountain lay elsewhere. Attempts to “optimize” any one hill in isolation merely displaced pain. This is the wickedness described in the first chapter: objectives couple across time and space, and measures become targets the moment livelihoods depend on them.
Escaping a local maximum requires no manifesto—only a safe, mechanical way to try better decisions and keep them when they pay. Experimental optimization provides exactly that: run side-by-side against end-of-day extracts; halt fast on insanity; expose drivers in coins; promote only on measured uplift; revert cheaply. Because the method prices trade-offs at firm scope, it dissolves many made-up “constraints” into penalties and turns departmental vetoes into tunable, auditable parameters.
In practice, the barrier is seldom physics; it is narrative. As long as averages and single KPIs narrate success, the equilibrium looks acceptable. Replace the narration with daily, auditable commitments priced in coins, and the supposed “constraints” reveal themselves for what they are: choices. The next section examines why the institutions meant to sharpen those choices—academia, vendors, consultants—so often teach averages and KPIs instead.
Why the equilibrium holds
Bad ideas need no shocks to survive. A supply chain can stay serviceable while its methods lag. The equilibrium persists because incentives are muted and misaligned: each actor involved in shaping the discipline earns a quiet rent from keeping decisions human-mediated and deniable, while no one is paid in coins for unattended decision quality.
Such equilibria endure because they clear a low bar. Trucks still move, rebates are captured, dashboards stay green, and no one must stake a career on a bolder alternative. In the absence of an existential shock, the local gains of inaction often outweigh the uncertain gains of rewiring decisions. As long as the apparatus works well enough, pressure to change remains muted.
The mechanics are banal. Long feedback loops dilute responsibility; dashboards substitute targets for economics; pilots are staged where failure cannot bite; “exceptions” divert the hardest cases back to humans, thereby insulating brittle logic from falsification. In such a setting, mediocrity compounds quietly. What passes for prudence—alerts before action, manual “validation”, consensus around a single forecast—merely preserves a clerical loop that neither prices uncertainty nor learns from it.
Around this loop sits a network of interests that damps pressure to change. Academia is rewarded for puzzles and publication, not for decisions that survive contact with docks and cut‑offs; it polishes stylized models that are easy to grade and hard to operationalize. Software vendors make their margins on seats, storage, and upgrades of systems of records and reports; genuine automation would shrink seats and erase screens. Consultants sell choreography—workshops, matrices, “alignment”—whose deliverables leave code and economics untouched; the engagement ends, the rituals remain, and the spreadsheet flow resumes. Corporate data-science teams, measured by projects and dashboards, stop at insights; ownership of the coin-denominated commitment sits elsewhere, along with the incentive to rework assumptions when reality disagrees. Inside the firm, supply chain remains second-class and subordinate to central IT: budgets favor ledgers that never fail over engines that must decide; risk is defined as platform deviation rather than economic loss.
Prices signal these incentives as clearly as any market does. Vendors bill for access, not uplift; consultants for effort, not gains; academics for novelty, not resilience; internal teams for artifacts, not emissions. Goodhart’s law then does the rest: once a metric promises career safety, the organization optimizes the metric. “Forecast accuracy” improves by shortening horizons and padding buffers; “service level” rises with inventory that nobody prices; “adherence to plan” is celebrated while trucks idle at a saturated dock.
The vignette above is typical, not exceptional. Nothing in physics prevented the weekly inbound pile-up; it persisted because the measures governing each silo—rebates, utilization, availability—were local, while the costs they displaced—demurrage, overtime, mid-week gaps—were borne elsewhere. Where no one is explicitly accountable for the firm‑wide economics of the decision, clerical equilibria harden into custom. The silence of profits and losses at the level of the micro‑choice leaves narratives—“best practice,” “industry standard,” “one set of numbers”—to fill the void.
The sections that follow examine the main sustainers of this equilibrium. Academia’s attachment to publishable puzzles, vendors’ rebranding of records and reports as “intelligence” while outsourcing intelligence to users through alert theater, consultants’ buzzworded ceremonies that stop short of code, and corporate data science’s insulation from ownership all contribute to a stable but mediocre state. The organizational trap—supply chain’s second-class status and subordination to central IT—locks the arrangement in place. None of this requires malice. It suffices that incentives are decoupled from unattended, auditable, coin‑denominated decisions. Until that coupling is restored, the equilibrium holds.
Who sustains it
The stalemate endures not through malice but through dispersed incentives. Four constituencies, each pursuing reasonable aims on its own terms, collectively sustain a stable but mediocre equilibrium: academia, enterprise software vendors, consultants, and corporate data‑science programs. Each constituency supplies something the others demand—publishable puzzles, marketable platforms, confidence-building ceremonies, and dashboards of “insights”—while leaving the daily act of allocation where it has long resided: with human clerks reconciling exceptions.
The lens adopted here is adversarial, not hostile. Earlier chapters have already set out what workable progress looks like: unattended, auditable emissions of decisions priced in coins. Against that yardstick, we can examine how each constituency, by following its own incentives, perpetuates practices that never clear the bar.
We begin with academia, whose stated mission—advancing knowledge—should have made it the natural antidote to stagnation. Instead, through a persistent misframing of supply chain, it has too often furnished the rhetoric of science without the substance of unattended decisions.
Academia
Universities and journals promise science; the factory floor receives spreadsheets. The gap is not a matter of taste but of framing. As argued in Epistemology, supply chain is applied economics: it prices options under variability to allocate scarce resources profitably. Recasting it as “applied mathematics”—a catalog of tidy puzzles with conveniently fixed assumptions—yields papers that read well and travel poorly. What matters operationally is not a theorem about an idealized reorder point; it is whether unattended software can turn records into auditable, coin-denominated commitments that survive contact with docks, cut-offs, rebates, and fat-tailed lead times.
This misframing explains why mainstream academic outputs so rarely register in practice. The canonical artifacts—point forecasts, Gaussian safety stocks tied to service targets, deterministic plans—suppress exactly what moves money: tails, correlations, batching, and ratchets that bind tomorrow’s options. When those artifacts are installed in software, they reappear as alert theater: the brittle logic delegates the difficult cases back to humans, who become the de facto intelligence. The organization then congratulates itself for “following best practice” while planners continue the clerical reconciliation computers were meant to end.
The puzzle filter compounds the problem. A publishable paper is easiest to produce when the world is simplified until a closed form or a small solver can be shown to “optimize” a toy. One more constraint is added here, one more index there; results are reported on synthetic instances or sanitized extracts. Yet nothing in this workflow compels the authors to expose their premises to falsification in the wild. In business, knowledge earns trust only when it places bets—emits decisions unattended—and improves profit. Absent emissions, the bar is not met, no matter how elegant the mathematics looks in isolation.
The neglect of data semantics and software engineering is equally telling. As detailed in Information, modern flows are mediated by overlapping systems of records and reports, where field meanings drift and where “derived data” (forecasts, ABC tags, buffers) contain no new information. A workable system of intelligence—the engine that issues unattended decisions—must therefore (i) read authoritative but messy records, (ii) carry uncertainty without flattening it, (iii) price trade‑offs in coins at firm scope, and (iv) write commitments back, with dossiers fit for audit. Most academic treatments airbrush this junction away. They assume clean, stationary inputs and stop where the work becomes decisive: semantics, economics, and code that runs every day at scale.
The right test bench has long been available. Intelligence, Decisions, and Deployment laid out a minimal discipline: dual-run against full-scope end-of-day extracts; halting heuristics that stop the engine when confidence drops; whiteboxed dossiers showing the monetary drivers of each line; promotion only on measured uplift; reversibility through statelessness and time-travel. A scientific contribution that claims operational merit should instantiate itself within that harness—so that failures surface as “insane lines” and are removed by revising costs, constraints, or semantics, not by exception lists and meetings. Papers that do not cross this threshold are chronicles or conjectures, not supply chain knowledge.
Curricula mirror the literature. Master’s programs continue to center on time‑series gadgets from the 1980s, Gaussian safety stocks tethered to arbitrary service targets, and classroom ABCs. Programming and data engineering are treated as accessories; enterprise semantics and probabilistic economics scarcely appear. Graduates learn to supervise spreadsheets and packaged “planning” modules, not to author numerical recipes that convert records into unattended emissions. The best student, sensing the ceiling, drifts toward other STEM fields where code, data, and falsification govern status. The remainder staff the clerical loop that the literature presupposes.
None of this condemns mathematics. Durable mathematical contributions do exist in our trade—but they present themselves as subroutines inside an economic and engineering frame: queueing results that calibrate congestion costs; linear‑programming components that assemble a truck; concentration bounds that regularize demand tails. The error is to mistake the subroutine for the system. An optimizer with the wrong objective, wrong semantics, or no unattended pathway is not an advance; it is a polished detour.
A workable academic agenda is close at hand. Treat supply chain as applied economics first; let mathematics, statistics, and computer science serve that end. Deliver artifacts that run: code that ingests semantically faithful extracts; price schedules in coins rather than “service targets”; probabilistic forecasts evaluated by proper scoring rules and uplift, not by retrospective fit; numerical recipes shipped whiteboxed so practitioners can audit each commitment. Above all, accept the discipline of experimental optimization: iterate in the field until insanity disappears, then publish what survived. Citations may follow; if not, the coins will.
Software vendors
Enterprise vendors sustain the equilibrium not by conspiracy but by economics and engineering fit. As argued in Information, there are three distinct software families: systems of records (authoritative ledgers and workflows), systems of reports (narrations of the past), and systems of intelligence (engines that issue unattended, auditable decisions). Most vendors sell records or reports, then repackage add-ons as “planning” or “optimization” to chase growth. The category confusion is not semantic; it is operational. A CRUD-first ledger cannot host heavy analytics without harming latency; a reporting stack built to reshape aggregates cannot carry uncertainty or price trade-offs in coins. When records or reports masquerade as intelligence, the flow reverts to clericalism.
The dominant pattern is alert theater. “Automation” takes the form of an exception list that delegates the hard cases back to humans. Users—recast as human coprocessors—manually reconcile thin rules against fat-tailed realities. Nothing in this posture improves the software’s intelligence; it merely hides the absence behind modern interfaces. The incentives align: vendors bill by seats, storage, and modules, not by uplift; genuine unattended decisions would reduce seats and screens. In Intelligence we noted the culture clash: records vendors optimize CRUD productivity and uptime; systems of intelligence optimize economic decision quality under uncertainty. The former has no incentive to build the latter.
Hype cycles amplify the misfit. Every few years, a wave of “AI‑driven planning”, “digital twins”, or “control towers” promises a clean break from spreadsheets. Startups are acquired and folded into suites; the add-on inherits the host’s record/report DNA and reappears as a configurable ruleset with alerts. The combinatorics remain unaddressed; variability is flattened into averages; costs are proxied by arbitrary service targets. The result is continuity under new acronyms.
Analyst quadrants stabilize the story. Because the research economy is funded by the vendors it rates, superficial features and fashionable labels are rewarded. The list of “leaders” rotates slowly; the criteria privilege breadth of catalog and referenceability over unattended decision quality. On the client side, RFP theater—critiqued in Deployment—produces adverse selection: questionnaires canonize yesterday’s artifacts, demos reward choreography, and the winner is the supplier best at answering questions that never should have been asked. When the stack goes live, the exception list returns.
Operationally, the failure modes are consistent. Category confusion creates resource contention (optimizers inside the ledger starve workflows) and information loss (derived artifacts are ingested as if they contained new information). Semantic drift—tolerable in a ledger when humans supervise—becomes lethal when fed to a decision engine that nobody lets decide. Where a system of intelligence would recompute from raw, authoritative records and write back priced, auditable commitments, the packaged “planning” tier consumes pre‑digested numbers and emits suggestions that planners must edit to stay safe.
None of this requires bad faith; it is the equilibrium implied by the business model. Vendors price access, not coins; buyers ask for screens, not unattended emissions. Thus, a stable mediocrity persists. The remedy is not a new quadrant but a different contract and a different architecture: keep records, reports, and intelligence strictly disjoint; run the decision engine as a stateless, replayable asset that ingests raw records, carries probabilistic inputs, prices trade‑offs in money, and writes commitments back with dossiers that expose their drivers. Select suppliers through adversarial market research rather than RFPs; pay in proportion to unattended decision scope and measured uplift. Where these conditions are met, alert theater dies of disuse; where they are rejected, spreadsheets resurface—an outcome that should be read as the diagnostic it is.
Consultants
Consultancies thrive by promising change that preserves what their clients are least willing to disturb: the applicative landscape, the org chart, and the daily rituals that assign accountability without owning outcomes. In the equilibrium described earlier, vendors sell records and reports under “planning” labels; consultants complete the loop by supplying the rhetoric and choreography that let brittle tools pass for progress. The deliverable is seldom a decision but a narrative: maturity models, quadrants, roadmaps, and ceremonies that rebrand yesterday’s routines.
The literature that underwrites these engagements is voluminous and shallow at once. It catalogs horizons, capabilities, and archetypes; it partitions the world into neat tiers and dial settings; it multiplies 2×2 matrices whose corners invite an infinity of pages. Such segmentations are not knowledge; they are staging. They evade the economic question—what allocation of scarce resources improves profit under variability—and substitute a taxonomy that can be taught and sold. As argued across this book, classifications earn their keep only when they predict; most consulting taxonomies do not.
Inside the engagement, buzzwords supply momentum. “S&OP alignment”, “integrated business planning”, “demand-driven”, “digital twin”, “control tower”, “resilience by design”—none is intrinsically absurd, but all are routinely deployed as covers for unchanged decision loops. The work product is workshops, cadence charts, and “one-number” committees. The same pattern reappears: exception lists grow, spreadsheets metastasize, and the hardest cases—fat tails, binders, ratchets, incoherent semantics—are politely routed back to the very clerks the software failed to relieve.
This cycle repeats because it is safe. Reorganizing meetings does not threaten the ERP; renaming phases does not expose premises to falsification; publishing a “target operating model” does not force a single coin-denominated commitment to be placed without human arbitration. The engagement pleases every constituency: executives receive a story of progress; middle management retains vetoes through ritual checkpoints; vendors keep modules in place; the consultant departs with references. The change in the flow of goods is negligible.
A workable standard exists to separate choreography from substance. In Information, Intelligence, Decisions, and Deployment we set the bar: daily, full‑scope dual‑run on authoritative extracts; promotion only on measured uplift in coins. By that yardstick, most consulting interventions are non‑events: they emit no unattended decisions, carry no probabilistic futures, price no trade‑offs, and accept no halting responsibility. They narrate; they do not allocate.
The economic test is simple. If a proposition cannot be run tomorrow night against the client’s end-of-day extracts, audited in coins, and—if it survives—paid for in proportion to uplift, it belongs to theater, not practice. When consultants are willing to sign for unattended emissions and accept reversibility and uplift as the basis for fees, they cease to be consultants in the conventional sense; they become co‑authors of the numerical recipe. This is rare, not because the work is mystical, but because the prevailing business model rewards ceremonies over code.
The “consultant cycle”—periodic transformations that rename the same routines—will persist until buyers invert the incentives. The remedy, already outlined in Deployment, is adversarial market research: have vendors and would‑be advisors draft and critique each other’s problem statements, demand back‑of‑the‑envelope TCOs and peer lists, and pay only for unattended scope and measured uplift. Where the engagement produces emissions, keep it; where it produces quadrants and cadences, let it lapse. The next section shows how a similar equilibrium reproduces itself inside the firm under the banner of “data science”—insights without ownership standing in for decisions.
Corporate data science
Inside firms, “data science” has largely reproduced the same equilibrium sustained by academia, vendors, and consultants: it produces insights without taking ownership of emissions. Since the early 2010s, teams have multiplied dashboards, pilots, and forecasts; yet few organizations can point to daily purchase orders, allocations, or prices issued unattended by code and written back to ledgers. The work looks modern, but it stops short of placing bets in the world.
Three traits recur. First, deliverables remain forecast‑centric and KPI‑centric. Error bars, accuracy scores, and tidy graphs abound, while the firm‑wide money ledger—the only common yardstick—is absent. Uncertainty is described, not priced. Second, the engineering junction is elided. Models are built atop derived artifacts (ABCs, buffers, “cleansed” tables) instead of authoritative records; field semantics drift; halting conditions are undefined. When a recommendation is brittle, an “exception” hands the hard case back to a planner. Third, incentives reward slides and proofs‑of‑concept rather than unattended commitments. “Decision support” becomes a convenient euphemism for preserving human arbitration over the very decisions software should make.
The result is insight theater that complements vendors’ alert theater. Planners end up reconciling brittle rules against fat-tailed realities by hand. The combinatorics that justify computers—millions of micro‑options each day—are pushed back to people. Clericalism, now animated by dashboards, persists.
The alternative is well within reach and has been sketched across this book. Replace the notion of an insight factory with a Supply Chain Scientist who owns a numerical recipe: a compact program that ingests authoritative records, carries probabilistic views of demand, lead times, and reliabilities, prices trade‑offs in coins at firm scope, and emits unattended decisions. The Scientist is accountable for emissions, not for decks. This posture forces tacit assumptions into the open and binds modeling to economics.
The operational vignette above (flattening the Monday pile‑up by pricing dock saturation and admitting “do nothing yet” as an option) is typical. No reorgs, no new warehouses: the same assets deliver better because decisions changed. This is what day‑to‑day ownership accomplishes. As soon as dock capacity carries a shadow price, MOQs become penalties rather than edicts, waiting competes economically with buying, and the weekly rhythm stops dictating losses. A system of intelligence turns such reasoning into unattended practice; a data‑science dashboard does not.
Accountability also imposes proportion. Once costs, constraints, and ratchets are encoded end‑to‑end, marginal gains in a sub‑model are pursued only when they move coins at the emission boundary. Fashionable metrics (forecast accuracy, service level) lose their talismanic status; what matters is uplift in risk‑adjusted operating profit after write-back. Collaboration then shifts naturally: finance publishes capital charges and valuations; procurement publishes rebates, MOQs, and reliabilities; operations publishes capacities and cutoffs. The recipe binds these inputs nightly under dual‑run, so learning compounds where it pays.
This is how the internal equilibrium breaks. Data science stops narrating the flow and starts running it—under safeguards that make automation safe to trust: daily replayable extracts, stateless computation, whiteboxed dossiers, and reversible emissions. The chapters Information, Intelligence, Decisions, and Deployment have already set the harness; here the point is institutional: until someone is paid for unattended decision quality, “data science” will continue to feed clerical loops. The next section arms readers with rejection filters for the literature that nourishes those loops—materials that look busy yet fail the tests of prediction, pricing, and falsification.
The organizational trap
External constituencies—academia, vendors, consultants—help explain why the equilibrium holds, but the decisive brake is internal to firms. Supply chain retains second-class status, de facto subordinated to corporate IT. Reputation, bureaucracy, and ticket latency interlock until the organization accepts clerical vigilance in lieu of engineered intelligence. The remedy is not another committee: a structural reallocation of ownership and a thin seam of data that frees flow from platform vetoes.
Second‑class status
Supply chain is praised as “mission‑critical”, yet treated as peripheral. Because trucks keep moving and shelves are stocked, leadership infers that only incremental polishing is warranted. The appearance of continuity becomes an alibi for underinvestment in code and the people who write it. A self-fulfilling loop follows: prestigious recruits aim elsewhere; the department inherits routines rather than authors decisions; shadow spreadsheets proliferate to compensate for fragile “planning” add-ons; the resulting clericalism confirms the original prejudice that nothing better is possible.
The reputation gap is old—logistics as the dull chore after the “real” work of product and brand—and it still shapes staffing and budgets. What makes the loop particularly tenacious is its intellectual veneer: singular, noneconomic KPIs — a “98 % service level,” “forecast accuracy” — masquerade as objective. When livelihoods depend on such surrogates, buffers inflate, exceptions are deferred to humans, and the ledger glows even as coins leak. The department looks busy and safe, even as it refuses the probabilistic economics that would measure trade‑offs in money and automate the mundane.
Bureaucracy and ticket latency
Where intelligence is manual, bureaucracy thrives. Alerts and exceptions are routed through rituals—monthly “one‑number” conclaves, pre‑allocation councils, approval cascades—whose true function is to distribute accountability rather than to improve decisions. Each ritual adds a layer of process without reducing uncertainty; the combinatorics remain unchanged, only now filtered through more hands.
Ticket latency completes the trap. Any substantive change—an added constraint, a revised cost schedule, a subtle semantic correction—requires filing tickets against central systems. Weeks turn to months as priorities are triaged by platform risk, not by expected economic uplift. By the time a ticket clears, the business has moved on and the patch is already stale. What should be a daily, reversible code edit inside a decision engine is recast as a cross-department project; the status quo prevails by default.
Compromise therefore drifts toward the smallest step all parties can tolerate. “Machine‑learning” modules are grafted on as dashboards that defer hard cases to an exception list; ERP upgrades unify fields but leave the decision rules untouched. The facade changes; the clerical loop persists.
Subordination to IT
Corporate IT’s charter is uptime, standardization, and compliance. This is proper for systems of records and reports. It is ill-suited to systems of intelligence. When “planning” is hosted inside the ledger, heavy analytics contend with workflows; when it sits atop reporting views, uncertainty is flattened and decisions devolve into alert theater—users as human coprocessors. Budgets follow the charter: capital flows to CRUD platforms; decision automation is purchased as a module and delivered as configuration plus exceptions.
Risk is inverted: deviation from the platform is treated as danger; economic loss from inferior decisions is treated as business as usual. Gatekeeping then hardens the boundary conditions: derived artifacts are treated as authoritative, schema changes are anathematized, and any tool that cannot be absorbed by the suite is rejected on principle. Under such governance, even sincere attempts at “advanced analytics” resolve into dashboards feeding the clerical loop.
Patch accretion
Half-measures accumulate. A temporary workbook to smooth a rebate ladder, a bespoke report that reconciles two masters, a per‑vendor exception in a buyer’s tool—each acquires dependents who rely on its outputs. Over time the patchwork gains defenders, and removing any one part appears riskier than leaving all in place. What began as a stopgap hardens into custom.
Patch accretion scatters accountability. No one owns the economic quality of the whole; each team owns a sliver and a story. This is wickedness in miniature: the problem mutates as each patch adapts to its neighbors, and survival is mistaken for success.
The cure is architectural rather than exhortatory. A seam that makes facts portable and a stateless recipe that turns those facts into priced, auditable commitments dissolve patches into parameters. With daily dual-run, veto power retreats; patches are either encoded properly or retired.
Seam and ownership
Two structural moves break the trap.
First, establish a seam: a faithful daily extract covering the systems of records. It is boring on purpose: append-only snapshots with manifests; enough history to cross a full cycle; no “helpful” upstream massaging; least-privilege reads. With it, the decision engine can be stateless and replayable. Tickets that once sought to alter ledgers become code changes against a private copy; semantics are audited empirically rather than negotiated by email. Latency shrinks from months to hours, and reversibility becomes routine.
Second, assign ownership of the numerical recipe to supply chain. Roles follow naturally (see Deployment): a Data Officer runs the seam; a Supply Chain Scientist authors the recipe; an Executive Sponsor arbitrates prices and boundaries; a Flow Manager brings field sense and escalates anomalies. Vendors and consultants are judged by emissions and measured uplift, not by screens.
Absent such a seam and such ownership, nobody is even positioned to ask—let alone answer—the only question that matters operationally: why are we not using modern computation to handle uncertainty and allocate scarce resources more profitably?
These two moves demote bureaucracy and platform vetoes from governors to parameters. Once a seam exists and ownership is clear, the department is no longer a second‑class custodian of other people’s software; it becomes the author of its economics in code. The rest of this chapter arms the reader with filters to reject materials that cannot survive such a harness. If a proposition cannot run nightly against the seam, and defend its emissions in coins, it belongs to theater, not practice.
Rejection filters
The bottleneck today is not a shortage of material but its surplus. More than a million papers and tens of thousands of books parade under the “supply chain” banner; the pile grows faster every year while day‑to‑day practice remains spreadsheet‑centric. The simplest explanation is that most of this production lacks the properties that make knowledge operational—properties that let unattended software place profitable, auditable bets every night. This section offers three filters to discard what will not help: trivia, authorities, and puzzles. They are not a taxonomy of the literature; they are a reader’s toolkit for protecting attention.
Trivia
Trivia catalogs the world—lists, matrices, horizons—without showing why the classification is essential, not accidental. Ockham’s razor applies: do not multiply categories beyond necessity. Here, necessity means power to predict and economize in coins. A segmentation earns a place only if learning it reduces uncertainty about future outcomes and improves the engine’s unattended allocation decisions.
The periodic table clarifies the standard: chemistry’s grid is not a decoration; it compresses causality. Atomic number orders the elements so behavior in reactions becomes predictable; the table is a compact, falsifiable theory of matter. By contrast, much planning jargon is mapmaking without geology. Splitting “planning” into strategic, tactical, and operational horizons says little unless the split predicts different money-denominated choices under uncertainty; otherwise the trichotomy is mere scenery. Likewise for product ABCs, “fast, slow, non‑moving”, “local, regional, national” sites, or the immortal 2×2 quadrants164 where “autocratic vs. consensus‑seeking” crosses “think‑then‑act vs. act‑then‑think”165. These devices are easy to teach and infinitely recombinable; their predictive content is typically nil.
The test is blunt: if adopting a classification does not change emissions—purchase orders, transfers, prices—it is trivia. In the earlier operational vignette, nothing about the weekly inbound pile-up required a new quadrant or horizon. The fix came from pricing dock saturation, replacing averages with distributions, and admitting “do nothing yet” whenever the option value of waiting dominated. A handful of prices and distributions moved coins; a shelf of segmentations would not.
Authorities
You’re in no position to lecture the public about anything. Opening monologue of the Golden Globes (2020), Ricky Gervais.
Prominence is not proof. Executive quotes, celebrity forewords, and brand‑name anecdotes entertain; they do not decide. Treat the appeal to authority as a red flag unless it crosses three bridges: falsifiability, counterfactuals, and emissions. Falsifiability means the claim could turn out wrong once encoded; counterfactuals require the speaker to state what would likely have happened otherwise; emissions require the claim to be precise enough to alter unattended commitments tomorrow night.
Large firms complicate the matter. Success attaches to many causes—product, brand, capital structure—while supply chain remains secretive by design. Even where the chain truly underpins success, the public spokesman may not be its author; outsized contributions are unevenly distributed, while many linger doing very little166. And those who are responsible have incentives to keep effective methods tacit or to recast them as palatable rituals. Counting quotes is not the road to knowledge.
The operational standard is narrow and fair. A proposition from a famed operator or vendor earns attention only once it can be written as code that ingests authoritative extracts at day’s close, emits, and defends each line in coins. Until then, file the story as a hypothesis and resist building processes around it. Authority may inspire; it may not substitute for falsification.
Puzzles
Academic supply‑chain literature tends to optimize what is tidy to write, grade, and publish—stylized puzzles with fixed assumptions. Puzzles can be delightful and, at times, useful as subroutines. They become harmful when presented as supply‑chain knowledge absent the junction where knowledge must live: semantics, economics, and code that runs every day. The issue is not mathematics; it is misframing. Supply chain is applied economics under variability; a paper that never places bets—never emits unattended decisions—is not yet in the arena.
Two questions separate a subroutine from a system. First, can the proposal ingest the original records without laundering meaning through derived artifacts? Second, does it carry uncertainty and price trade-offs all the way to the emission boundary? If the answer to either is “no”, the contribution remains a puzzle. The correct harness has been described throughout this book: daily dual‑run on full‑scope extracts; halting heuristics; dossiers that expose the top monetary drivers of each line; promotion only on measured uplift; reversibility through stateless, replayable computation. A method that cannot survive this harness is not “early stage”—it is untested.
This filter is not hostile to theory; it is hospitable to reality. Queueing results that calibrate congestion costs, programming blocks that assemble a truck, concentration bounds that regularize fat tails—these are welcome once embedded in a numerical recipe that moves coins. Mathematical elegance without unattended emissions is mere storytelling in symbols. The literature will modernize when papers ship code that passes the same economic test we require of production: under the same constraints and audit trail, does it raise risk-adjusted operating profit?
Using the filters
Applied consistently, these filters compress a reading list to materials that change emissions. Trivia falls to Ockham’s razor unless it demonstrates predictive, money-denominated power; authorities are accepted only once their claims can be falsified under dual-run; puzzles advance only when they attach to unattended, auditable decisions. This is why so much activity coexists with little progress: most of what circulates fails these tests while looking busy. The remedy is not cynicism but discipline: price uncertainty, demand emissions, and let coins—not credentials—arbitrate.
Looking onward
Looking onward, the point is not another manifesto but a change of posture. The yardstick proposed throughout this book is practical and narrow: a supply chain advances when, every night at the closing bell, software emits unattended, auditable commitments priced in coins—or halts itself and explains why. Everything else is staging. By this measure, progress is neither a quarterly ceremony nor a new screen; it is the steady replacement of clerical reconciliation with a mechanical flow that carries uncertainty explicitly and prices trade-offs at the scale of the firm.
The path requires no revelation—only two durable moves that place responsibility where it can compound. First, make facts portable through a seam: schema‑faithful, immutable extracts that let the decision engine be stateless and replayable. Second, condense the firm’s economics into a numerical recipe that ingests those extracts, carries probabilistic views of the future, assigns penalties and rewards in money, and emits unattended decisions. Whiteboxed dossiers that accompany each emission expose the few drivers that mattered in coins. With this discipline, improvement becomes routine: when reality falsifies an assumption, the recipe changes, replays, and resumes. Errors do not accumulate; knowledge does.
People do not vanish; their work acquires leverage. Planners become stewards of frames and prices rather than line‑editors of suggestions. Finance publishes capital charges and valuations that the engine must honor; procurement publishes rebates, MOQs, and reliabilities as prices rather than vetoes; operations publishes capacities and cut-offs as auditable constraints. Leadership arbitrates trade-offs in money, not in slogans, and owns the rare occasions when the engine halts. The test is the same for everyone: did the emissions raise the firm’s risk-adjusted rate of return?
Alternatives—alert theater, plan theater, insight theater—can be kept alive indefinitely because trucks still move and averages still comfort. Its cost is largely invisible in a day’s snapshot yet crushing over years: options neglected, tails mispriced, attention spent grooming artifacts that carry no information. The choice facing mature firms is therefore not between stability and risk, but between a clerical equilibrium that quietly taxes every flow and a mechanical discipline that makes uncertainty pay.
Before the constructive program, we must be blunt about what stands in the way. The next subsection names the habits that guarantee clerical equilibria no matter the software or the rhetoric. Only after discarding them does the better path—already sketched in earlier chapters and entirely compatible with existing ledgers—become straightforward to adopt.
What must stop
If progress is to be more than a slogan, three entrenched performances must cease. Each has a pedigree; each flatters professionalism; and each reliably preserves clerical equilibria while computers idle.
First, alert theater must stop. Software that styles itself “automated” yet emits lists of exceptions for humans to reconcile is not automation; it is an interface for postponing decisions. A system that cannot write back unattended and expose its drivers in coins reduces planners to reconciling exceptions by hand—and calls it progress. The criterion is simple and non‑negotiable: either the program issues commitments to the ledgers—orders, allocations, prices—or it is a reporting tool and should be judged (and purchased) as such. Anything that requires daily line‑editing to be “safe” belongs to the past. Seats, screens, and alerts are no substitute for emissions.
Second, plan theater must stop. Monthly consensus rituals around a “one set of numbers” flatten uncertainty, with ignorance masquerading as precision. Forecast accuracies, service‑level targets, and other plan‑adherence indicators are not economics; they are numerically phrased etiquette. They reward buffers, not judgment; compliance, not profit. Where edicts are declared “hard constraints”, they typically mask unpriced preferences; where averages stand in for futures, tails are ignored until they bite. The remedy is not decorum but prices: treat most limits as penalized constraints in coins; let “do nothing yet” compete with “commit now” through an explicit option value of waiting; replace plan adherence with the only test that matters—under the same constraints and audit trail, did the emissions raise the firm’s risk‑adjusted rate of return? Absent this discipline, plans will continue to look impeccable while trucks idle at saturated docks.
Third, insight theater must stop. Dashboards, control towers, and “digital twins” that never place a bet are ornaments. So are pilots that live forever in sandboxes, “visibility‑only” programs that move no levers, and procurement rites (RFPs) that select vendors on choreography rather than unattended decision scope. Inside the firm, the mirror image is a data-science function that produces slides and prototypes yet refuses ownership of write-backs. The standard is again mechanical: nightly dual‑run on faithful extracts; stateless, replayable computation; halting on insanity; promotion only on measured uplift; reversibility by design. Anything less narrates the flow; it does not contribute to it.
These three performances—alerts in lieu of decisions, plans in lieu of prices, insights in lieu of emissions—persist because they are comfortable and deniable. They end the moment leadership demands coin-denominated unattended decisions rather than seats and ceremonies. Only then does the constructive program that follows have room to work at all.
Towards a better way
The better way is prosaic in routine, radical in consequence. It treats tomorrow’s uncertainty as priced raw material and turns the daily close into a disciplined reckoning. From authoritative records, a compact numerical recipe recomputes the firm’s admissible moves, carries probabilistic views of the future, attaches coin-denominated payoffs and penalties, and either writes back unattended commitments or halts itself when confidence drops. Nothing else qualifies as progress. Everything that matters—options, tails, couplings, ratchets—is expressed in code that can be audited, replayed, and revised without ceremony.
Once this discipline is in place, the day takes on a different texture. Insanity disappears first: dock saturation is no longer a surprise averaged away in weekly dashboards; “MOQs as edicts” dissolve into priced trade-offs; an explicit option value lets “do nothing yet” compete fairly with “commit now”. Trust accumulates not in meetings but through whiteboxed dossiers that explain, line by line, in coins, why an order was placed, a transfer withheld, or a price nudged. When reality falsifies an assumption, the engine halts; the premise is corrected; the run replays; the flow resumes. Wrongness becomes cheap and reversible; rightness scales.
Two brief vignettes show this in practice. A seasonal fashion label opens with generous bets but replaces push plans with policy: assortments are evaluated as options; early sell-through updates a basket-level probabilistic view; markdowns become a reified function rather than after-the-fact remorse. The engine prices stockout pain against obsolescence in one ledger, lets late binding prevail where half-life is long, and writes unattended allocations and price moves each night. The visible effects are not theatrical: fewer mid-season transfers, smaller end-of-season fire sales, and steadier gross margin. Beneath the surface, the gains come from admitting tails, pricing them, and letting postponement win when it should.
In civil aviation MRO, a rotable pool is governed as a portfolio of options. Unscheduled removals are modeled as fat-tailed arrivals; AOG exposure carries a shadow price; cannibalization, lease, repair, and buy are admissible moves ranked by expected risk‑adjusted return. Many nights the most profitable action is to wait; on others, the engine pre-positions stock against a rising risk shadow, not a calendar. What used to be a chain of emails becomes unattended dispatch and purchase orders. Each carries a dossier that exposes, in coins, why a part flew to this tail number, why another stayed on the shelf, and why a lease beat a repair given turnaround dispersion. Fewer crises are “resolved” by heroics because fewer are manufactured by earlier edicts.
Institutional consequences follow naturally. Roles tilt from arbitration to authorship: planners become stewards of frames and prices; finance owns capital charges and valuations the recipe must honor; procurement publishes rebates, MOQs, and reliabilities as schedules rather than vetoes; operations states capacities and cut-offs as auditable constraints. Vendors and advisors are paid for unattended scope and measured uplift, not for seats, screens, or ceremonies. Budgets move toward the only asset that compounds this century: a small codebase that writes the firm’s economics down and improves under experimental optimization.
None of this demands revelation. The instruments have been laid out already to achieve unattended decisions. Demand a change of posture: accept that every commitment is a bet whose quality can be stated in coins ex ante and audited ex post, and organize work so that software places these bets at the cadence reality allows.
The book opened by insisting that supply chain is an intent, not a thing, and closes by making that intent executable. When a firm can explain each nightly commitment in coins and stop itself when it cannot, it has crossed the line that separates theater from mastery. The rest is slow compounding of knowledge—seeing a little farther each evening and foreseeing enough to deserve tomorrow’s return.
Technical notes
Aperiodic rate of return
This annex formalizes the aperiodic rate of return referenced in Chapter Economics. The intent is practical: each option (a micro‑decision that commits scarce resources) should be ranked by the speed at which its tied‑up capital is repaid and begins to compound. Because supply‑chain cash flows seldom align with calendar periods, the option itself selects its most informative horizon.
We consider a discrete time grid $$t=0,1,\dots,T_{\texttt{max}}$$ (days, weeks, or months). Let
$$ s(t)\in\mathbb{R}^+ \quad\text{and}\quad r(t)\in\mathbb{R}^+ $$
denote, respectively, the present‑valued spending and revenue at time $$t$$ associated with the option. (If discounting is desired, convert future amounts to present value first; see the next section.)
Define the total capital the option requires—valued at $$t=0$$—and the cumulative revenue up to any cutoff $$T$$:
$$ S ;\triangleq; \sum_{t=0}^{T_{\texttt{max}}} s(t), \qquad R(T) ;\triangleq; \sum_{t=0}^{T} r(t). $$
The aperiodic rate of return is then
$$ \mathtt{RoR}{\texttt{aper}} ;=; \max{0<T\le T_{\texttt{max}}} \frac{1}{T}\Bigl(\frac{R(T)}{S}-1\Bigr). $$
Interpretation. Slide a cursor along the option’s cash‑flow timeline. At each date $$T$$, compute the per‑unit‑time return you would have realized if you had closed the position then: cumulative inflow $$R(T)$$, divided by the total capital $$S$$ the option demands, minus one, scaled by $$1/T$$. The metric keeps the peak value; the corresponding $$T^\star$$ is the option’s implicit horizon. Two options with very different calendars become commensurable because each is judged at its own best repayment point.
Units. If $$t$$ is measured in days, $$\mathtt{RoR}_{\texttt{aper}}$$ is a daily rate. Annualization—when desired—follows compounding on the chosen grid (e.g., $$(1+\rho)^{365}-1$$ for a daily rate $$\rho$$). For small $$\rho$$, multiplying by the number of periods is a coarse approximation; compounding is safer once rates are large.
Worked example. A buy today of 100 coins ($$S=100$$) yields two sell‑throughs: 60 coins at day 15 and 60 coins at day 30. The per‑day speeds are
$$ \rho(15)=\frac{1}{15}\Bigl(\frac{60}{100}-1\Bigr)=-0.026\bar{6},\quad \rho(30)=\frac{1}{30}\Bigl(\frac{120}{100}-1\Bigr)=0.006\bar{6}. $$
The maximum occurs at $$T^\star=30$$ days with $$\rho^\star\approx 0.006\bar{6}$$ per day; compounded over a year this is roughly $$(1+0.006\bar{6})^{365}-1\approx 10.3$$ (i.e., about $$1{,}030%$$ per annum). Such magnitudes are commonplace at the item level because they reflect fast inventory rotations rather than portfolio‑wide returns.
Two caveats. First, a positive bias: the metric assumes that the released capital at $$T^\star$$ can be redeployed immediately into an equally attractive opportunity. In practice, matching opportunities may arrive with a delay. Second, a negative bias: revenues beyond $$T^\star$$ are ignored for the sake of measuring speed. In reality, unless goods are scrapped, later inflows will still materialize and lift the ultimate profit. Both effects are acceptable—indeed useful—because the purpose here is not to settle the final P&L of the option, but to prioritize scarce capital toward the fastest compounding uses.
Discounted cash flows
When we discount cash flows, we merely restate the earlier discussion of time preference in coin and on a discrete time grid. A coin received later is worth less today because it cannot be used, buffered, or reinvested in the meantime. Discounted cash flow (DCF) converts every future amount into “today‑coins” and, once this conversion is done, the aperiodic rate of return of the previous section can be applied without any further change.
Fix a time grid $$t=0,1,\dots,T_{\texttt{max}}$$ (days, weeks, or months). Let $$k$$ denote the firm’s cost of capital per grid step; if the cost of funds varies with time (credit line tiers, temporary squeezes), write $$k_t$$. The discount factor that brings an amount at date $$t$$ back to $$t=0$$ is
$$ D(t);=;\prod_{\tau=1}^{t}\frac{1}{1+k_\tau} \qquad\text{(for constant $k$, $D(t)=(1+k)^{-t}$).} $$
Let $$r^{}(t)$$ and $$s^{}(t)$$ be the undiscounted revenue and spending at time $$t$$ (with immediate amounts recorded at $$t=0$$). Their present‑valued counterparts are
$$ r(t);=;D(t),r^{}(t), \qquad s(t);=;D(t),s^{}(t). $$
The discounted cash flow of the whole option (the sequence it entails) is then
$$ \mathtt{DCF} ;=; \sum_{t=0}^{T_{\texttt{max}}} \bigl(r(t)-s(t)\bigr) ;=; \sum_{t=0}^{T_{\texttt{max}}} D(t),\bigl(r^{}(t)-s^{}(t)\bigr). $$
This sum is an absolute valuation in today‑coins (when positive, it creates wealth; when negative: it destroys value).
Connection to the aperiodic rate of return. In the previous section, $$r(t)$$ and $$s(t)$$ were explicitly defined as present‑valued flows. Thus, to rank options by speed, first discount as above, then plug the present‑valued series into the aperiodic rate formula unchanged. In symbols, keep $$S=\sum_{t=0}^{T_{\texttt{max}}} s(t)$$ and $$R(T)=\sum_{t=0}^{T} r(t)$$, and evaluate
$$ \mathtt{RoR}{\texttt{aper}} ;=; \max{0<T\le T_{\texttt{max}}}; \frac{1}{T}\Bigl(\frac{R(T)}{S}-1\Bigr). $$
Discounting enforces time preference; the aperiodic rate then reports how fast the present‑valued outlay is repaid and begins to compound.
Choosing $$k$$. Use the firm’s marginal cost of liquidity on the grid you compute with. If you work with an annual rate $$K$$ on a daily grid, convert it to a per‑day step by $$k=(1+K)^{1/365}-1$$ (mutatis mutandis for weeks or months). When liquidity is piecewise or time‑varying, encode it as a path $${k_t}$$; the definition of $$D(t)$$ already accommodates this. In finance it is common to “pad” $$k$$ to reflect uncertainty; in supply chain we avoid this double counting—uncertainty is handled explicitly via the probabilistic treatment of returns later in the annex.
Worked micro‑example (daily grid). Spend 100 coins at $$t=0$$; receive 60 coins at day 15 and 60 coins at day 30. With an annual cost of capital of 10%, the daily step is $$k\approx(1.10)^{1/365}-1$$. The discount factors are $$D(15)\approx (1.10)^{-15/365}$$ and $$D(30)\approx (1.10)^{-30/365}$$. The DCF is
$$ -100 ;+; 60,D(15) ;+; 60,D(30) ;\approx; 12.6\ \text{coins.} $$
Feeding the present‑valued series into the aperiodic metric gives $$S=100$$ and $$R(30)\approx 60,D(15)+60,D(30)\approx 112.6$$, hence
$$ \mathtt{RoR}_{\texttt{aper}};\approx;\frac{1}{30}\Bigl(\frac{112.6}{100}-1\Bigr);\approx;0.0042\ \text{per day}, $$
which compounds to roughly $$(1+0.0042)^{365}-1\approx 2.6$$ (about $$260%$$ per annum). The precise figure is not the point; the method is. Discount first to express everything in today‑coins; then compare options by how fast those present‑valued inflows repay the present‑valued outlay.
Practical notes. Units must match: $$t$$ is measured in the same step as $$k$$. If payments are staged (e.g., a deposit at $$t=0$$ and a balance at shipment), treat each as its own $$s^{*}(t)$$. While interest rates matter, inventory rotations typically dominate in supply chain; discounting’s role here is to restore comparability across options with different calendars so the aperiodic rate can do its ranking job cleanly.
Informational entropy
This annex distills the few pieces of Shannon’s theory used in Chapter Information, in a form directly usable for inventory examples. We stay with discrete variables and base‑2 logarithms so units are bits (changing the log base merely rescales the unit).
Consider a discrete random variable $$X$$ that takes states $$x$$ with probabilities $$p(x)$$. The information revealed when we learn that the state is $$x$$—the self‑information—is
$$ \operatorname{I}(x);=;-\log_{2} p(x)\quad\text{bits}. $$
A certain outcome ($$p=1$$) carries $$0$$ bit; rarer outcomes carry more bits.
The entropy of $$X$$ is the average self‑information:
$$ \mathrm{H}(X);=;\mathbb{E}[\operatorname{I}(X)];=;-\sum_{x} p(x),\log_{2} p(x)\quad\text{bits}. $$
Entropy depends only on the probabilities, not on how we name or store the states. It can be read as the average number of fair yes/no questions needed to identify the state. Two immediate corollaries suffice for our purposes:
-
If $$X$$ has $$n$$ equally likely states, then $$\mathrm{H}(X)=\log_{2} n$$.
-
If $$X_1,\dots,X_k$$ are independent, the joint uncertainty adds:
$$ \mathrm{H}(X_1,\dots,X_k);=;\sum_{j=1}^{k}\mathrm{H}(X_j). $$
(When variables are correlated the joint entropy is below that sum; we will assume independence here to keep the illustration simple.)
We will model each product’s stock level as a discrete variable and, as a first pass, treat different products as independent. This lets us compute a store’s entropy as “100 times” the per‑product value. The next subsection applies these formulas to two contrasting stores—a luxury watch shop and a suburban depot—and makes the contrast quantitative. For a compact, accessible primer on the wider theory, see James V. Stone, Information Theory: A Tutorial Introduction (2019).
Entropy of two stores
We now make precise the contrast used in the original chapter. Model each SKU’s on‑hand as a discrete random variable and, as a first pass, treat SKUs as independent.167 The store‑level entropy is then the sum of the per‑SKU entropies.
Luxury watch shop. Each SKU has at most one unit on hand, with a 95% probability of being available and a 5% probability of being out of stock. Thus a single SKU is binary with
$$ p(1)=0.95,\qquad p(0)=0.05. $$
Using the binary entropy,
$$ \mathrm{H}{\text{watch}}^{(1)} ;=; -,0.95,\log{2}(0.95);-;0.05,\log_{2}(0.05);\approx;0.2864\ \text{bits}. $$
For 100 independent SKUs,
$$ \mathrm{H}{\text{watch}} ;=; 100,\mathrm{H}{\text{watch}}^{(1)} ;\approx; 28.64\ \text{bits}. $$
Suburban depot. For each SKU, assume the on‑hand can be any integer from 0 to 9,999 with equal probability. A uniform law on $$n$$ states has entropy $$\log_{2} n$$, hence per SKU
$$ \mathrm{H}{\text{depot}}^{(1)} ;=; \log{2}(10{,}000);\approx; 13.2877\ \text{bits}, $$
and for 100 SKUs
$$ \mathrm{H}{\text{depot}} ;=; 100,\mathrm{H}{\text{depot}}^{(1)} ;\approx; 1328.77\ \text{bits}. $$
Numerically, the depot carries about $$\mathrm{H}{\text{depot}}/\mathrm{H}{\text{watch}} \approx 46.4$$ times more inventory information than the watch shop.
Read in “twenty‑questions” terms, fully specifying the watch shop’s entire 100‑SKU stock state takes on average about 29 yes/no answers; the depot takes about 1,329. The uniform assumption slightly overstates entropy relative to realistic, peaked stock distributions, but the gap remains of the same order. This calculation underwrites the chapter’s claim that two assortments of identical size (100 SKUs) can have radically different informational weight—one measured in a few dozen bits, the other in kilobits—hence the necessity of software for anything beyond the simplest flows.
Topological sort
A topological sort arranges a collection of items so that every prerequisite appears before the item that depends on it. The input is a directed graph whose nodes are items (tasks, tables, components) and whose arrows encode “must come before”. If the graph has no directed cycle, such an order exists; if there is a cycle, no ordering can honor all prerequisites.
Supply‑chain practice is full of such precedence structures. A bill of materials requires sub‑assemblies before the assembly; a daily numerical recipe must ingest records before deriving features, forecast before valuing options, and value options before emitting purchase orders or transfers. In a system of intelligence, this structure is made explicit as a directed acyclic graph (DAG) and executed in that order at each run.
Intuition. Start with everything that has no prerequisite; place it in a worklist $$S$$. Repeatedly pick one item $$n$$ from $$S$$, output it, and relax its outgoing arrows: each successor $$m$$ has one fewer prerequisite; when that count reaches zero, $$m$$ joins $$S$$. If, in the end, some items were never output, they mutually depend on one another—a cycle.
Worked supply‑chain example. Consider a tiny bill of materials: Screw and Panel have no prerequisites; Frame requires Panel; Assembly requires Screw and Frame. One valid order is:
$$ \text{Screw},\ \text{Panel},\ \text{Frame},\ \text{Assembly}. $$
Running the algorithm, $$S$$ initially holds Screw and Panel; after outputting Panel, the in‑degree of Frame drops to zero and it joins $$S$$; after Screw and Frame have been output, Assembly can be produced. The exact order among simultaneously eligible items is irrelevant to correctness but matters for reproducibility.
Correctness and complexity. Every edge $$(u!\to!v)$$ is honored because $$v$$ can only be output after its last prerequisite is removed, which happens only after $$u$$ has been output. Conversely, if an item is output it had no unmet prerequisite at that moment. The method runs in linear time $$O(|V|+|E|)$$ with adjacency lists and a single pass over edges, and uses linear space. In production systems, make outputs deterministic by taking items from $$S$$ in a fixed key order (e.g., $$(\texttt{site},\texttt{sku},\texttt{step})$$); the same inputs then reproduce bit‑for‑bit the same order, which supports auditing and time‑travel replays.
Operational use in a system of intelligence. The daily numerical recipe is a DAG: ingest raw extracts, normalize, derive features, fit or refresh probabilistic models, compute valuations in coins, then emit idempotent write‑backs (POs, transfers, prices). The runtime performs a topological sort of that computation graph and executes nodes accordingly. A detected cycle is not a corner case to be “overridden”; it is a breach of semantics that must halt the engine (as argued in Chapter Intelligence) and surface a small witness cycle to fix the logic or the data contract. Producing such a witness is straightforward: after the algorithm stops with $$|L|<|V|$$, follow predecessor links within the residual subgraph until a node repeats; the loop thus revealed is the minimal explanation the operator needs.
Four-layer perceptron
A four‑layer multilayer perceptron (MLP) is the smallest useful specimen of what the “deep learning” section described in words: a fixed circuit made of linear maps separated by simple non‑linear gates. It is a good vehicle for demystifying the mechanics because every step reduces to ordinary linear algebra.
We revisit the handwritten‑digit toy task (MNIST) used earlier; see Figure \ref{fig:perceptron} in Chapter Intelligence. Each gray‑scale image is a $$28\times 28$$ array rescaled to numbers in $$[0,1]$$ and flattened into a vector $$\mathbf{x}\in\mathbb{R}^{784}$$. The network computes scores for the ten digits $$0,\dots,9$$, then turns those scores into probabilities. The predicted digit is the one with the highest probability.
Forward pass (inference)
The network stacks three affine transforms separated by rectifiers. Using column‑vectors and applying non‑linearities element‑wise, the computation is:
$$ \begin{aligned} \mathbf{z}_1 &= \mathbf{W}_1,\mathbf{x} + \mathbf{b}_1, & \quad \mathbf{a}_1 &= R(\mathbf{z}_1),\ \mathbf{z}_2 &= \mathbf{W}_2,\mathbf{a}_1 + \mathbf{b}_2, & \quad \mathbf{a}_2 &= R(\mathbf{z}_2),\ \mathbf{y} &= \mathbf{W}_3,\mathbf{a}_2 + \mathbf{b}_3. && \end{aligned} $$
Here $$\mathbf{W}_1!\in!\mathbb{R}^{256\times 784}$$, $$\mathbf{W}_2!\in!\mathbb{R}^{128\times 256}$$, $$\mathbf{W}_3!\in!\mathbb{R}^{10\times 128}$$ are weight matrices; $$\mathbf{b}_1!\in!\mathbb{R}^{256}$$, $$\mathbf{b}_2!\in!\mathbb{R}^{128}$$, $$\mathbf{b}_3!\in!\mathbb{R}^{10}$$ are bias vectors. The non‑linearity is the rectifier
$$ R(u);=;\max(0,u), $$
applied coordinate‑wise. Rectifiers are not only sufficient for expressiveness; they are also cheap to evaluate on silicon, which is why they displaced older sigmoid‑shaped gates.
The output $$\mathbf{y}\in\mathbb{R}^{10}$$ collects one score per digit. To obtain class probabilities we apply the soft‑max map:
$$ p_k ;=; \frac{e^{y_k}}{\sum_{j=0}^{9} e^{y_j}} \quad\text{for}\quad k=0,\dots,9. $$
The classifier reports $$\arg\max_k p_k$$; the full probability vector $$\mathbf{p}$$ is useful during training because it quantifies uncertainty and supports a smooth loss.
Training by stochastic gradient descent
Training chooses all weights and biases so that the network assigns high probability to the correct digit on images it has not seen. Let $$(\mathbf{x}^{(i)},,t^{(i)})$$ denote the $$i$$‑th labeled example, with $$t^{(i)}\in{0,\dots,9}$$ the true class. The standard loss for a mini‑batch $$B$$ is the average cross‑entropy
$$ \mathcal{L}B ;=; -,\frac{1}{|B|}\sum{i\in B}\log p_{,t^{(i)}}(\mathbf{x}^{(i)}), $$
optionally augmented by a small quadratic penalty on the weights (weight decay) to curb overfitting. Modern frameworks compute the gradient of $$\mathcal{L}_B$$ with respect to every parameter by automatic differentiation. One step of stochastic gradient descent (SGD) then nudges parameters in the direction that reduces the loss:
$$ \theta \leftarrow \theta ;-; \eta,\nabla_{\theta},\mathcal{L}_B, $$
where $$\theta$$ denotes the collection of all $$\mathbf{W}\ell,\mathbf{b}\ell$$ and $$\eta>0$$ is the learning‑rate. Repeating this simple loop—sample a mini‑batch, run the forward pass, evaluate the loss, apply the gradient update—gradually lowers the loss on held‑out images until improvements stall. Variants (momentum, adaptive learning rates) refine the same idea; none alter the basic picture.
Two practical choices matter for stability and speed but do not complicate the story. First, initialize weights so that activations neither blow up nor vanish as they propagate: with rectifiers, a zero‑mean Gaussian scaled like $$,\mathcal{N}(0,;2/n_{\text{in}})$$ works well, $$n_{\text{in}}$$ being the fan‑in of the layer. Second, process examples in small contiguous groups (“mini‑batches”) to exploit hardware parallelism; inference and training then reduce to a handful of dense matrix products (GEMMs) plus cheap rectifiers.
Sizes, costs, and what is really computed
Counting only the weights, the $$784!\to!256!\to!128!\to!10$$ MLP carries
$$ 784\times256 ;+; 256\times128 ;+; 128\times10 ;=; 234,752 $$
trainable numbers; the three bias vectors add a negligible $$256+128+10$$ on top. One forward pass is “three matrix products and two rectifiers.” This mechanical view—linear algebra punctuated by a simple gate—is the correct mental model. There is no hidden magic; the network is a programmable circuit whose parameters are tuned by gradient descent.
Mathematical optimization
This annex complements the discussion The optimization paradigm in Chapter Decisions. Its purpose is practical: to restate the vocabulary of that section—marginal commitments, feasibility tests, priced penalties, probabilistic forecasts, and economic objectives—in a minimal mathematical form that admits unattended, repeated resolution.
From decisions to a model
A decision is a finite set of marginal commitments. Let $$\mathcal{A}={a_1,\dots,a_n}$$ denote the admissible micro‑moves for the present cycle (buy one more unit, allocate one more unit to store $$s$$, schedule one more stop, and so on). Selecting a move is recorded by $$x_i\in{0,1}$$; the vector $$x\in{0,1}^n$$ encodes the whole commitment. The frame makes $$\mathcal{A}$$ explicit and attaches to each $$a_i$$ the data needed to test feasibility and to value consequences.
Feasibility is expressed as tests that a candidate $$x$$ must pass. Some limits are absolute—legal prohibitions, physical impossibilities—and appear as hard tests that reject $$x$$ outright. Most operational limits tolerate relaxation at a price; rather than forbidding those violations we price them. Let $$g_k(x,\omega)$$ be a measurable quantity the firm wishes to restrain (dock load on a day, missed window minutes, overtime hours), computed under a realization $$\omega$$ of the uncertain future. A priced penalty $$C_k(g_k)$$, stated in coins, turns that preference into the objective. Keeping only truly hard limits as tests keeps the engine flexible and honest; priced penalties make unlike nuisances commensurable on one ledger.
The objective is money‑denominated. Within the window of responsibility attached to the present decision class, each micro‑move $$a_i$$ earns coin flows (margins, rebates, salvage) and incurs flows (procurement, handling, carrying, obsolescence, congestion, late fees). Under a probabilistic view of futures $$\Omega$$, the engine evaluates the expected value of $$x$$,
$$ J(x);=;\mathbb{E}\Omega\Bigg[;\sum{i=1}^{n} x_i,v_i(\Omega);-;\sum_{k} C_k\bigl(g_k(x,\Omega)\bigr);\Bigg], $$
where $$v_i(\Omega)$$ is the net coin contribution of $$a_i$$ over the window once all ordinary costs are charged. Risk adjustments—e.g., heavier pricing of tail losses—live in the $$C_k$$ terms rather than as abstract percentages. When ranking small alternatives that tie up capital for different durations, Chapter Economics recommends comparing by aperiodic rate of return; in practice, this means computing both the expected coin delta and the tied‑up capital and then reporting speed with the metric defined earlier in this annex. The solver can maximize $$J$$ and, for emitted marginals, report the associated aperiodic speeds.
Uncertainty is carried in the inputs, not hidden in the objective. The random element $$\Omega$$ collects the distributions the firm maintains for demand, lead times, reliabilities, returns, and any priced exogenous factor. The objective $$J$$ is deterministic given those distributions. This separation keeps causality and blame legible when outcomes disappoint: either the frame omitted options, or prices were misguided, or forecasts were miscalibrated.
Penalties versus chance bounds
Two formalisms encode “keep this under control”. A chance constraint requires $$\mathbb{P}\bigl(g_k(x,\Omega)\leqslant \tau_k\bigr)\geqslant 1-\alpha_k$$ for some threshold $$\tau_k$$ and tolerance $$\alpha_k$$. A priced constraint assigns a coin cost $$C_k(g_k)$$ to all values of $$g_k$$. The former hides economics inside a dimensionless $$\alpha_k$$ that must be tuned elsewhere; the latter states trade‑offs directly in money and aggregates naturally with other terms. In supply‑chain operations—where most limits can be overrun at a cost—the priced form is the workhorse; pure chance bounds are best kept as guardrails for rare, safety‑critical fronts.
A convenient choice is a convex, rising penalty, often piecewise linear, that approximates the firm’s nuisance curve. For dock load $$q$$ beyond a soft capacity $$K$$ with sharp pain after $$K+\Delta$$, one writes $$C(q)=c_1,(q-K)+ + c_2,(q-K-\Delta)+$$ with $$c_2\gg c_1$$. The exact numbers are economics; once stated, the solver finds least‑bad compromises across heterogeneous frictions.
Worked micro‑model: filling a container
An importer must raise a full‑container order. Each candidate unit $$a_i$$ carries a volume $$w_i$$ and an expected, window‑bounded net coin $$m_i(\Omega)$$ that already includes procurement, handling, holding, obsolescence, salvage, and shortage effects. The container capacity is $$W$$ in the same unit as $$w_i$$. Overfilling is possible but painful at rate $$\lambda$$ per excess unit.
A faithful objective is
$$ \max_{x\in{0,1}^n};\mathbb{E}\Omega\biggl[\sum_i x_i,m_i(\Omega);-;\lambda\Bigl(\sum_i x_i,w_i - W\Bigr)+\biggr], $$
subject to hard tests such as legal incompatibilities or forbidden co‑loadings. This is a penalized knapsack under uncertainty. A simple, effective resolution aligns with the chapter’s marginal stance: at each step compute a shadow price for capacity, pick the admissible unit with the highest expected marginal $$\frac{m_i}{w_i}$$ above that shadow price, recompute the shadows, and stop when the next unit fails the economics of waiting cut‑off (expected, risk‑adjusted aperiodic return below capital plus option value of delay). Because valuation depends on the set in the box and not on the ceremony of its selection, this greedy accumulation targets what matters economically; the window and terminal valuations ensure clean attribution.
Worked micro‑model: smoothing inbound
A warehouse faces random supplier drifts and a finite number of receiving doors. Let $$d_t(x,\Omega)$$ be the number of arrivals on day $$t$$ implied by $$x$$ under $$\Omega$$. Overtime, driver detention, demurrage, and disruption are captured by a convex penalty $$C_{\text{dock}}(d_t)$$. A daily, coin‑denominated objective over horizon $$T$$ is
$$ \max_{x};\mathbb{E}\Omega\Bigg[\sum_i x_i,m_i(\Omega);-;\sum{t=1}^{T} C_{\text{dock}}\bigl(d_t(x,\Omega)\bigr)\Bigg], $$
again with only truly hard limits enforced as tests. The optimizer advances or defers orders until the expected return of the next admissible batch exceeds the shadow cost of capital plus the option value of waiting, while the convex $$C_{\text{dock}}$$ spreads load without brittle edicts.
-
Jomini derives it from the French logis, “lodgings.” ↩︎
-
By the 2020s, operations research (OR)—in Americanized spelling—remains productive, but mostly in designing solvers for mathematical optimization, with little connection to business or supply chain practice. ↩︎
-
Patented in 1952 by Norman Woodland and Bernard Silver. ↩︎
-
Also invented in the 1980s. ↩︎
-
This insight is incorrect – deeply so – but we will return to it at a later time. ↩︎
-
Here “intelligence” means collecting and presenting data; by any reasonable standard those tools are not intelligent. ↩︎
-
The 2020–2021 shortage of next‑generation game consoles illustrated this dynamic: resale prices on major platforms frequently reached double the manufacturer’s suggested retail price. ↩︎
-
As of June 2024, Google Scholar returns 827,000 papers for “inventory optimization techniques”. While there are certainly duplicates and false positives, one million published papers is the correct order of magnitude. ↩︎
-
See “The Logic of Scientific Discovery” (1934) by Karl Popper, a landmark in the history of science. ↩︎
-
In contrast, Austrian economists were entirely unsurprised by those outcomes. It was exactly what their theory predicted. ↩︎
-
The “anomalous” orbit of Mercury around the sun was instrumental in rejecting the old Newtonian mechanics in favor of the relativity theory of Einstein. In fact, there was no anomaly, merely an incorrect theory. Once relativity was adopted, the anomaly vanished. ↩︎
-
Amazon.com search performed in 2025; the overwhelming majority of results are authored by consultants rather than academics. ↩︎
-
Two millennia ago, Caesar had employed this exact approach with his Commentarii de Bello Gallico (Gallic War), going as far as claiming the Romans suffered no deaths in very large battles; a feat considered as radically implausible by historians. ↩︎
-
Projecting a great deal of confidence, competence, and sympathy is second nature for those who rise through the ranks of large corporations. They are essentially prerequisites for corporate careers. ↩︎
-
Stack ranking, also called forced ranking or “rank-and-yank”, is a performance review method where employees are placed on a bell curve—from top performers at one end to low performers at the other—compelling managers to “reward the best” and “punish the worst,” regardless of the absolute performance level of each individual. ↩︎
-
As already observed by Thomas Sowell in Knowledge and Decisions (1980) and Intellectuals and Society (2010), as well as by Ludwig von Mises in Human Action (1949), political commentary from economists frequently stems from ideology rather than genuine economic analysis. The author is hardly the first to point out this phenomenon. ↩︎
-
One of the most basic aspects of managing a conflict of interest is that the person subject to the conflicting position doesn’t get to decide whether his views are biased or not. The conflict of interest is an a priori cause to assume severe bias. Moreover, considering the considerable role that 21^st^-century states play in the economy at large, proposing that an economist funded by the state apparatus isn’t massively conflicted is quite an extravagant proposition. ↩︎
-
Some products are mass produced in a way that ensures that every unit is visually unique, typically through a pseudo-random coloring or patterning technique. However, those variations are intentional and carefully controlled. These products are not an exception to the standardization principle. On the contrary, they reflect an even tighter mastery of production processes, as they rely on an intentional variability without compromising any other product qualities. ↩︎
-
Enterprise software products are frequently sluggish and unresponsive. At the time of this writing, it remains frequent for an operation triggered by a user to take many seconds to complete. However, considering that even entry-level computing systems now feature gigabytes of memory, gigaflops of processing, and gigabits per second of bandwidth, such events must almost always be attributed to the incompetence of certain software vendors, not to any inherent limitations of the computing hardware with respect to the complexity of the operation requested by the user. ↩︎
-
Government interventions in the economy lead to crony capitalism, where corporate success depends on collecting handouts rather than satisfying customers. The term “profit” ought therefore to be reserved for gains achieved in an unhampered market. Wealth acquired through coercion—even when the coercer is a government—differs categorically from the surplus earned through voluntary exchange, although conventional P&L statements blur the distinction. ↩︎
-
In France, for decades, the slogan of left-leaning parties has been “Nos vies valent plus que leurs profits” (our lives are worth more than their profits). ↩︎
-
For example, from an inventory replenishment perspective, the method referred to as “ABC analysis” proposes to organize products in 3 to 5 classes depending on their respective volumes. Each class is assigned its own target service level used to dimension the corresponding safety stocks. The ABC/XYZ is a variant that segments products based on projected demand and projected erraticity. Those noneconomic methods betray a lack of understanding of the nature of supply chain. ↩︎
-
The sharp fall in U.S. freight rates after the deregulation of trucking in 1980 (Motor Carrier Act of 1980, Public Law 96-296.) shows how quickly efficiency returns once artificial moats are removed. ↩︎
-
Mercantilism, which was the dominant economic view from the 16th to the 18^th^ century in Europe, presents countries as economic adversaries. Exports are seen as favorable, while imports are seen as unfavorable. Local businesses are frequently eager to promote a mercantile perspective as it is a proven method to have the political apparatus block, or at least hinder, foreign competition. However, while countries can be at war in a geopolitical sense, there is no such thing as “economic warfare”. Free markets are based on voluntary transactions that only happen if both parties view the transaction as beneficial. Mercantilism is another long-disproven economic perspective that still remains popular to the present day. ↩︎
-
The fate of the company and the fate of its employees are distinct, and employees are fully aware of that. Even if a loyalty program is in place through stock options or similar means, getting an internal promotion is almost invariably more impactful for the employee than whatever reward the loyalty program might offer. If an employee has to choose between a promotion and the reward of the loyalty program, it is relatively safe to bet that the promotion will be chosen. ↩︎
-
For an extensive survey of disastrous ideas that enjoyed massive popularity among intellectuals in their time, see Intellectuals and Society (2010) by Thomas Sowell. ↩︎
-
Anticapitalism has been dominant among intellectuals for the last 150 years or so. The pursuit of profits is seen as one of the most serious defects of capitalism. The non-profit status supposedly fixes this defect and thus must be rewarded with some sort of tax exemption and/or regulatory relief with respect to obligations imposed on regular for-profit businesses. Whether the pursuit of profits is generally good or bad for society as a whole is a concern of praxeology, the science of human action, not of supply chain. For an extended discussion, see Human Action (1949) by Ludwig von Mises. ↩︎
-
Associations are frequently funded, at least in part, through taxation in the form of government subsidies. This only makes the government one of their customers. Associations routinely have enterprise customers as well. The specifics of the customers are a technicality that has no bearing on the general economic operating principles governing an association. ↩︎
-
Those rules invariably neglect spending on research and development, hence underestimating the break-even point, and they invariably neglect revenues from complementary products or enhanced brand awareness, hence overestimating it. In short, those rules are nonsense. ↩︎
-
In this context, “goodwill” refers to the common meaning—friendly, helpful, or cooperative attitudes—rather than the accounting definition of an intangible asset. The shadow valuations discussed here are inherently linked to decision-making. ↩︎
-
Even for a component serving a single finished product, pitfalls remain: the unavailability of a low-value part may jeopardize the flow of a high-value product. ↩︎
-
The effectiveness of the free market has long baffled intellectuals. The idea that an unintelligent mechanism—ad hoc voluntary exchanges—can produce better—smarter, even—solutions than those experts devise is profoundly surprising. Despite the resounding successes of free markets over the last three centuries, and the disasters of planned economies, most intellectuals remain skeptical. See also The Anti-Capitalistic Mentality (1956) by Ludwig von Mises. ↩︎
-
These binary digits are commonly called bits. We will later reuse the word “bit” in Shannon’s information‑theoretic sense; context will make it clear whether we mean a binary digit or a unit of information. ↩︎
-
As of August 2024, some 1 TB solid-state drives weigh about 8 grams and sell for roughly the cost of four Paris haircuts. ↩︎
-
Some countries regulate how long personal data can be stored. Unless a supply chain software vendor mishandles personal data within its system, these regulations generally have little impact on supply chain operations. Personal data offers limited value for optimizing the flow of goods and is more of a liability than an asset in this context. ↩︎
-
Event sourcing requires all events affecting the system to be stored in an event store, replacing the more common relational database. Unlike relational databases, where unlogged state changes can occur, event sourcing ensures by design that every change is logged, capturing all information that led to the system’s current state. ↩︎
-
Since mobile devices far outnumber other consumer computing devices, making apps resource-efficient also helps preserve battery life, a highly valued feature for users. While storing more data doesn’t inherently consume more energy, the more data that must be stored, retrieved, and processed, the more energy the device will use. ↩︎
-
Perishable goods, which have expiration dates, and serialized inventory, where each item has a unique serial number, are two notable exceptions to the usual SKU perspective. In these cases, stock on hand alone doesn’t adequately represent the inventory, so a more complex model is needed. ↩︎
-
By mind, we refer to an entity capable of general intelligence, putting aside all the concerns of consciousness and soul which have no bearing on supply chains. Such entities were once exclusively the province of man; however, recent advances in software and hardware are increasingly blurring that line. ↩︎
-
This 13-page essay is very well written and largely accessible to a general audience without any particular prerequisite. Free copies can easily be found online; we strongly recommend this essay to the curious reader. ↩︎
-
Classical MRP explosions, MOQ/EOQ “optimizations”, and time‑bucketed reorder cycles embedded in ERPs are deterministic, parameter‑driven recipes. They neither price optionality nor balance opportunity costs across the firm; hence they do not qualify as planning under the definition advanced in this book. ↩︎
-
A more recent architectural pattern, event sourcing, records every state change as an immutable stream of events instead of persisting only the latest state. Event‑sourced systems deliver superior traceability, replayability, and auditability, yet the pattern has been fully commoditized since the mid‑2010s: mainstream databases, frameworks, and cloud services all provide off‑the‑shelf support. Consequently, every conclusion drawn in this subsection for CRUD—about commoditization, resource contention, and cultural mismatch with systems of intelligence—applies verbatim to event‑sourced systems as well. ↩︎
-
Upselling also works with subscriptions. In this case, when the vendor sells an extra, the subscription fee is increased. ↩︎
-
Colloquially: the supposedly “single source of truth” for core business entities such as products, locations, suppliers, and customers. In practice, there is almost never a single source, only competing partial authorities. Large “MDM” suites promise a golden record through central hubs, survivorship rules, and stewardship workflows. In practice, they often become yet another overlapped ledger whose governance drifts as fast as the others. ↩︎
-
Hollow-core optical fibers (HCF) guide light mostly through air rather than silica, bringing propagation speeds closer to $$c$$ and yielding, on long links, single-digit to low-double-digit percentage latency cuts. Even if HCF were widespread, switching, buffering, routing detours, and protocol handshakes would keep end-to-end decision latencies largely bounded; the macro-argument—that latencies are now a hard constraint for enterprise systems—still stands. ↩︎
-
Command Query Responsibility Segregation. ↩︎
-
In the 2010s, fashion e-commerce in Germany routinely featured return rates above 50% of the units ordered. Indeed, many leading e-commerce retailers had generous return policies, and patrons were used to order multiple sizes for their garments and return all the units that didn’t fit. ↩︎
-
If counting in-house developments, then the ‘e-commerce front end’ app has been redeveloped independently more than a thousand times worldwide since the early 2000s. ↩︎
-
Historically, shadow systems have almost never been systems of intelligence. True decision engines demand durable data integrations, empirically validated modeling, version-controlled code, resource isolation, and runtime monitoring—requirements that exceed what spreadsheets or ad hoc desktop databases can deliver. Consequently, shadow efforts concentrated on records and reports. However, since the early 2020s, large language models and agent frameworks have made it feasible for non-IT teams to assemble narrow, quasi-autonomous decision bots. This “shadow intelligence” will dominantly arise in small, well-scoped niches (e.g., survey and qualify potential suppliers based on public web data) where data plumbing requirements are minimal. ↩︎
-
From the mid-1990s to the late 2010s, the database management system Microsoft Access enjoyed considerable popularity in large companies, allowing a white-collar employee to roll out his own system of records backed by a relational structure. ↩︎
-
Perfect numerical identity would require mirroring the source system’s floating-point quirks, rounding rules, and version-specific bugs. For decision-making, economic equivalence—not byte-for-byte equality—is what matters; vanishing numerical differences do not change the ranking of options nor the resulting profitability. ↩︎
-
An arms race happened in the 2010s between publishers—possibly operators of large marketplaces—and scrapers, the former trying to prevent the latter. However, for all practical intents and purposes, through the use of headless browsers, proxy networks, and more recently LLMs (large language models), the scrapers emerged victorious. Considering the evolution of the software technologies, there is little reason to ever expect any lasting reversal of this situation. ↩︎
-
Competitive intelligence tools also frequently reveal that competitors themselves occasionally face pricing glitches, with a low percentage of their own prices being bogus, either due to some “fat finger” clerical error, or due to bugs in the e-commerce front end itself. ↩︎
-
Archiving raw pages or API payloads is indispensable for audits and parser regression tests. It also guards against silent upstream changes. Observe the target site’s terms of use and applicable law; rate-limit scrapers and respect robots.txt directives unless a specific legal basis and business necessity justify otherwise. ↩︎
-
A click farm is a setup where a large group of low-paid workers or automated systems are employed to artificially inflate online metrics, such as clicks, likes, views, or app downloads, to manipulate the popularity or ranking of digital content, apps, or social media accounts. ↩︎
-
Full automation is also a practical requirement for a dual-run—operating the new decision engine in parallel with the legacy process—without duplicating the planning workforce. We return to this point later. ↩︎
-
In any case, those systems were heavily promoted as such by their respective software vendors. ↩︎
-
It became the ASCM (Association for Supply Chain Management) in 2018. ↩︎
-
In 2023, the Linux kernel has over 30 million lines of code, while Microsoft Windows has over 50 million lines of code. Even Android, a mobile operating system, has more than 7 million lines of code when counting dependencies. ↩︎
-
At a minimum, those companies include the corporate clients of Lokad. Furthermore, while technological giants, like Amazon or JD.com, don’t fully publicize the fine print of their technologies, their research publications on the same time period suggest they have also uncovered methods capable of delivering very high levels of automation, quite comparable to those achieved by Lokad. ↩︎
-
Mainstream ERP/APS offerings still delegate decisions to human clerks and function primarily as systems of records or reports. ↩︎
-
Numerous examples of such companies can be found in history. For example, Kodak was a pioneer in photography and had developed many patents related to digital cameras. Kodak employee Steven Sasson developed the first handheld digital camera in 1975. Larry Matteson, another employee, wrote a report in 1979 predicting a complete shift to digital photography would occur by 2010. Yet they hesitated to embrace digital photography, fearing cannibalization of film. That reluctance hastened their decline and culminated in a 2012 bankruptcy as competitors seized the digital market. ↩︎
-
For example, cloud computing was popularized by Amazon, deep learning by Google. ↩︎
-
Modern relational engines offer extensive mitigations; nevertheless, engineering complex operations under a constant resource budget remains challenging. ↩︎
-
A theory of mind reflects the understanding that others’ beliefs, desires, intentions, emotions, and thoughts may be different from one’s own. ↩︎
-
Mind-enhancing drugs are a favorite science-fiction trope—Dune’s Sapho juice, for instance—but no such known compound rivals anabolic steroids’ impact on muscle. ↩︎
-
Silver et al., Mastering the Game of Go without Human Knowledge (2017). AlphaGo Zero trained against itself and beat its predecessor 100-0. ↩︎
-
Bubeck et al., Sparks of Artificial General Intelligence: Early experiments with GPT-4 (2023). ↩︎
-
The vehicle routing problem is known to be an NP-hard problem. In computer science, NP-hard problems are a class of problems that are jointly conjectured to be strictly more difficult to solve than any given polynomial-time algorithm. Exhibiting a polynomial-time solution for any NP-hard problem would refute the conjecture, proving that all NP-hard problems can, in fact, be solved in polynomial time. ↩︎
-
One can also do this unintelligently by inventing arbitrary variables and constraints, then applying standard algorithms—a pattern common in academic supply-chain papers. ↩︎
-
ASCM, founded in 1957 as APICS, has issued over 100,000 CPIM certifications since 1973. ↩︎
-
Titles vary—demand planner, inventory analyst, replenishment manager, scheduler, and so on. ↩︎
-
Multimodal models first convert images into a latent vector that occupies token slots, so no explicit image-to-token step is needed. ↩︎
-
In computer science, a tensor is the dimensional generalization of the matrix. A scalar, a vector, and a matrix are respectively tensors of dimensions zero, one, and two. ↩︎
-
A feat performed through differentiable programming, which combines automatic differentiation with stochastic gradient descent. ↩︎
-
“Prompt engineering” sounds grand; most bright people learn those skills in a few days. ↩︎
-
Tool names and wire formats keep changing, but the pattern is stable: the model proposes, the runtime executes, state returns to the model. ↩︎
-
See Large Language Models are Zero-Shot Reasoners (2023) by Kojima et al. ↩︎
-
ChatGPT by OpenAI was released in November 2022. ↩︎
-
Operations research was originally understood as research on (military) operations. ↩︎
-
As of September 2024, a Google Scholar query for “inventory optimization” returns 1.9 million publications. It is hardly an exaggeration to say that none of this literature has ever survived in real-world conditions. ↩︎
-
Techniques such as cutting-plane methods, branch-and-bound, and branch-and-cut exemplify the mainstream view that a solver must not only provide good solutions but also prove that the solutions are either optimal or quasi-optimal. ↩︎
-
While nowadays learning problems and optimization problems have accumulated sharply distinct collections of techniques, it is still not entirely clear whether this distinction is entirely appropriate. Indeed, most learning techniques do have an optimization algorithm at their core, and conversely, some authors—including the present one—believe that the future of optimization lies in introducing a learning component at its core. In the future, these two fields might merge if a suitable unifying paradigm is discovered. ↩︎
-
Expert systems are collections of logical and arithmetic rules written by domain experts. They were largely abandoned in the 2000s, as they generally proved more costly and less efficient than learning-based alternatives. However, rule-based subsystems are still prevalent in enterprise software, and many have reached an operational complexity well beyond what was considered “advanced” in the early 1990s. ↩︎
-
L.G. Valiant’s 1984 paper A Theory of the Learnable introduced the concept of probably approximately correct (PAC) learning, laying the first rigorous foundation—later called computational learning theory—for studying learning from a statistical standpoint. PAC theory, however, ultimately proved less influential than the subsequent Vapnik–Chervonenkis framework. ↩︎
-
See Support-Vector Networks (1995) by Corina Cortes and Vladimir Vapnik. ↩︎
-
See Greedy Function Approximation: A Gradient Boosting Machine (1999) by Jerome Friedman. ↩︎
-
See Random Forests (2001) by Leo Breiman. ↩︎
-
In everyday experience with two or three dimensions, shapes such as circles and spheres fill a significant fraction of the enclosing squares and cubes. In higher dimensions, however, the volume of a hypersphere becomes a vanishingly small fraction of the hypercube’s volume; almost all the space lies in the corners, well beyond the reach of the centrally placed hypersphere. ↩︎
-
Technically, cross-entropy computed with fixed-precision floating-point arithmetic does have a definite minimum, since the range of representable values is finite. However, such a minimum is largely meaningless because numerical overflows and underflows cause critical failures long before this threshold is reached. ↩︎
-
Fuzzy logic—a form of many-valued logic in which variables can take any real number between 0 and 1—was originally developed in the 1960s. It might be considered the first programmatic paradigm for what was then called “artificial intelligence”, although it never became more than an extremely narrow niche with limited applications. ↩︎
-
The scikit-learn open-source project, initiated by French data scientist David Cournapeau in 2007, exemplifies this approach. ↩︎
-
In machine learning, “labels” refer to known instances or known solutions to a given learning task. For example, in automated translation from French to English, the label for a French text is its corresponding, accurately translated English text. ↩︎
-
See xVal: A Continuous Number Encoding for Large Language Models (2023) by Golkar et al. ↩︎
-
While chemistry has advanced to where it is common knowledge that “air” is 78% nitrogen and 21% oxygen, with the remainder mostly argon and water vapor plus traces of other gases, it still takes a deliberate effort to bring those facts to mind. ↩︎
-
Ancient Greek philosophers like Aristotle classified air as one of the four fundamental elements—alongside earth, water, and fire—without recognizing its composite nature. Only in the late 18^th^ century, through scientists such as Antoine Laurent de Lavoisier, was air’s true composition established. ↩︎
-
Standard Oil, General Electric, Ford, Kodak, IKEA, McDonald’s, Apple are but a few that became, at some point, vision-led giants. ↩︎
-
The Oxford dictionary defines “teleological” as “relating to or involving the explanation of phenomena in terms of the purpose they serve rather than of the cause by which they arise.” ↩︎
-
See Intellectuals and Society (2009) by Thomas Sowell. ↩︎
-
The modern flavor of weather forecasting predates economic forecasting by more than half a century, dating back to the invention of the electric telegraph in 1835. By the late 1840s, weather forecasts could be made from knowledge of weather conditions farther upwind, because information traveled faster than the winds themselves. However, weather forecasting belongs to the natural sciences, whereas economic forecasting does not. ↩︎
-
Newton’s third law states that when two bodies interact, they apply forces to one another that are equal in magnitude and opposite in direction. ↩︎
-
A time series featuring equal time intervals is said to be equispaced. In practice, non-equispaced time series are so rare in the literature that any reference to a time series implicitly assumes equal intervals. However, note that monthly and yearly intervals are not strictly “equal”: both months and years have varying numbers of days. ↩︎
-
Nearly all fundamental laws of physics, such as Maxwell’s equations, are time-reversal symmetric. Reversing time ($$t \mapsto -t$$) leaves their equations unchanged. However, the Standard Model exhibits time-reversal violation via the weak force (CP violation). TCP violation was discovered and experimentally verified in the 1960s. Nevertheless, it remains an advanced consideration, largely ignored outside physics—especially at the “vision” level. ↩︎
-
A Google Scholar survey over roughly the past 50 years (query run in 2024) indicates a ratio on the order of 1,000 papers mentioning “demand forecasting” for every paper mentioning “lead time forecasting.” The imbalance is striking given that nearly every firm operating a supply chain must anticipate both future demand and future lead times. ↩︎
-
This perspective is exemplified by the book Flowcasting the Retail Supply Chain (2006) by American supply chain consultants Mike Doherty and Andre Martin. The flowcasting approach consists of establishing daily forecasts at the store level, SKU by SKU, and of deriving all the allocations across the multiple echelons of the network from those “foundational” forecasts. ↩︎
-
Most countries nowadays have all sorts of compelled transactions in place, forcing people to pay for goods or services they didn’t choose: schooling, healthcare, roads, vaccines, foreign aid, etc. Assessing whether those compelled transactions are a net positive for society or not, and in which circumstances they can be considered exploitative, is beyond the present scope. Let’s however point out that those “transactions” have nothing to do with a free market. ↩︎
-
The “price gouging” laws are the archetype of this flawed understanding of economics in general and of supply chains in particular. Those laws restrict both the capacity of buyers to get what they want, when they want it the most—thus, incurring a net loss for those underserved buyers—but also restrict the capacity for the sellers to expand their offering through extra investments that could only be recouped if their prices are higher than usual. ↩︎
-
See The Anti-Capitalistic Mentality (1956) by Ludwig von Mises for an extensive discussion on why intellectual elites, and academia in particular, are almost invariably fierce opponents of entrepreneurial behaviors, which includes the rugged vision presented in this chapter. ↩︎
-
In the aviation industry, an “AOG” (aircraft on ground) refers to an incident where an aircraft is grounded due to a missing part or any other technical issue that must be addressed prior to letting the aircraft fly again. Such incidents are costly: a single delay can cascade into missed connections and, if extended, force the carrier to provide hotel rooms for stranded passengers. ↩︎
-
Large bureaucracies tend to evolve mechanisms that diffuse accountability—layers of committees, ambiguous mandates, and gameable metrics. A textbook case is the Space Shuttle Challenger accident (1986): the Rogers Commission concluded that the launch decision was structurally flawed, with responsibility fragmented across NASA and contractor management and dissent filtered out by process rather than owned by any single decision-maker. For the broader pattern, compare Parkinson’s Law (1955) and Pournelle’s “Iron Law of Bureaucracy” (2010). ↩︎
-
These intellectual circles also include management consultants who, like academics, depend on a steady stream of publications to gain the notoriety their careers demand. ↩︎
-
See Wal-Mart Stores, Inc. (1994), European Case Clearing House, by Stephen P. Bradley, Pankaj Ghemawat, and Sharon Foley. ↩︎
-
When visiting a local grocery store, all it takes to undermine the forecast of the retailer consists of emptying the shelf of a given product, immediately causing a stockout. While a few products are stored in large quantities making this exercise impractical, such as bottled water, many more products are stored in very limited quantities, low single digit quantities, and thus, making it quite affordable to adopt such an adversarial behavior. ↩︎
-
On Google Scholar (November 2024), the query “time series forecasting” returns roughly 4,230,000 results. The figure is indicative only: duplicates and near-duplicates inflate it, while missed items deflate it. Either way, a quick audit confirms that the correct order of magnitude is indeed “millions”. ↩︎
-
In classical mechanics, momentum equals mass times velocity. It is a vector quantity. ↩︎
-
Black Friday and its derivatives (e.g., Cyber Monday) are defined relative to civic holidays, not by astronomical cycles. ↩︎
-
Ramadan migrates roughly 10–11 days earlier each Gregorian year; Chinese New Year falls between late January and mid-February. Use dedicated binary calendars for such events rather than smearing their effects across months. ↩︎
-
Our argument only applies to supply chain settings. Natural sciences offer all sort of periodic measurements which can also be approached through the time-series paradigm. Those time series may exhibit all sort of sophisticated patterns that can be used to anticipate future events with great accuracy. Conversely, in pure finance, when no physical goods are involved, even a minuscule pattern can be relentlessly exploited by an actor to arbitrage the market and generate profits in the process. In contrast, in supply chain, statistical patterns need to be nontrivial to be harnessed. ↩︎
-
See The M5 Uncertainty competition: Results, findings and conclusions (2020), by Spyros Makridakis, Evangelos Spiliotis, Vassilis Assimakopoulos, Zhi Chen. In the overall score, which involved forecasts performed over a dozen of aggregation levels, the Lokad team was initially ranked No. 6, but this ranking was later revised to No. 5, as one of the better-ranked teams was found guilty of cheating the competition. ↩︎
-
See A white-boxed ISSM approach to estimate uncertainty distributions of Walmart sales (2021) by Rafael de Rezende, Katharina Egert, Ignacio Marin, Guilherme Thompson, December 2021 ↩︎
-
A steady trend over multiple years can be seen as a particular case of momentum, where the business is expected to keep growing as it did in the recent past. Introducing a trend factor in a time-series model is straightforward, and numerous techniques exist to do so. ↩︎
-
New time-series forecasting models have been introduced steadily for seven decades. Dominant trends include parametric models (1950–1980), hidden Markov models (1990s), support vector machines (2000s), gradient-boosted machines (2000s), deep learning (2010s), and generative pretrained large models (2020s). ↩︎
-
Strictly speaking, if there is a nonzero probability that delivery never occurs (e.g., loss, destruction, seizure), the mean lead time is infinite: some probability mass sits at “never”, so the expected lead time diverges. In practice, medians and quantiles are more meaningful than means for such distributions. ↩︎
-
1,000 units per day, or 0 units per day, or something else entirely. Indeed, the customer can also simply change its ordering volume; we ignore that third option here for clarity, though real settings would require considering it. ↩︎
-
The mode of a probabilistic distribution is the point where the probability mass function attains its maximum. The mode need not be unique. ↩︎
-
Wealth’s fat-tailed inequality is not peculiar to capitalism. Feudalism, socialism, and other authoritarian regimes yield even greater inequality than capitalism. Laws and customs do not change the nature of a phenomenon—specifically, whether it is thin- or fat-tailed. Naturally, this has never stopped anyone from promising the opposite to his fellow men. ↩︎
-
On March 12, 2020, sales of toilet paper were up 845% week-over-week versus the prior month, according to NCSolutions, which aggregates U.S. retail purchase data. ↩︎
-
A telltale misuse is modeling lead times as normal. Beyond thin tails, a Gaussian assigns nonzero probability to negative delays—implying shipments arrive before they are sent. Any model that lets time run backward is unfit for operations. ↩︎
-
A “market study” is the bureaucrat’s comfort blanket: slow, expensive, and low-signal. It produces slides and committees, not knowledge and decisions. Accountability diffuses to the point of vanishing while time-to-learning stretches by quarters. If the goal is to learn, small, instrumented field trials beat purchased reports almost every time. ↩︎
-
If the products were indeed perfect substitutes, then the store manager would have no reason to keep both products as part of the store’s assortment. Let’s suspend our disbelief for a second, a more convincing, yet more complicated example will be provided afterward. ↩︎
-
“Supply Chain Digital Twin” is a marketing term many vendors use for their simulation offerings. ↩︎
-
The “Simple Sabotage Field Manual” (1944), published by the Office of Strategic Services (precursor of the CIA), recommends bureaucratic sabotage. The infiltrated agent must insist that the organization apply all policies, all rules, and all regulations to the letter (no matter how trivial), thereby burying the organization in red tape and slowing progress to a crawl. Far from being an insight of historical interest, modern companies operating large supply chains are more susceptible to bureaucratic sabotage than ever—so much so that accidental bureaucratic self-sabotage has become part of the daily routine for many, if not most, large companies. ↩︎
-
An IBM presentation from 1979 ↩︎
-
For example, Amazon’s Supply-Chain Optimization Technologies (SCOT) team has end-to-end responsibility for orchestrating Amazon Store’s supply chain—demand forecasting, buying, inventory placement, and customer-promise decisions. Senior leadership decided in 2014 to make several product categories “100% automated”, with subsequent rollout across the retail business. ↩︎
-
Warehouses, once requiring sizable workforces, have been increasingly replaced by fully automated warehouses since the 2010s. Transportation will follow. In 2024, there are over 2 million fully self-driving vehicles in operation worldwide. ↩︎
-
An algorithm is superlinear when its time or memory rises faster than the problem size; doubling the variables can more than double the cost. ↩︎
-
This separation was canonized by modeling languages such as AMPL: A Modeling Language for Mathematical Programming (1993), Fourer et al. The model/solve split is pedagogically clean and excellent for benchmarking; in day-to-day supply-chain optimization it often impedes the “mechanical sympathy” that practice demands—warm starts from yesterday’s plan, domain-aware decompositions, cached subcomputations, and explicit economic instrumentation. ↩︎
-
This contrast is not accidental but the product of selection. Industrial supply chains are continuously pruned by market pressure: configurations that require brittle, all-or-nothing “perfect plans” fail disproportionately and are retired; what remains is a population of routines, contracts, and physical layouts that tolerate noise, delays, and partial failure. Robustness is seldom the result of explicit design alone; it accumulates through survivorship. As a consequence, most operational choices admit neighborhoods of near-equivalent moves that deliver comparable returns, which is why approximate, greedy search guided by correct prices often outperforms the pursuit of exact global optima in toy models. ↩︎
-
Thomas S. Kuhn, The Structure of Scientific Revolutions (1962). In business settings, paradigm shifts are typically small and frequent: a new admissible move, a repriced penalty, a different cadence. Their effect on the “best” plan is nevertheless incommensurable with what came before. ↩︎
-
Gene Amdahl, 1967: the speedup of a system is bounded by the fraction that cannot be parallelized. In decision engines the non‑parallelizable parts are often I/O, semantic checks, or global reconciliations that keep the solution coherent. ↩︎
-
“Detention” compensates a carrier when a truck waits beyond free time at loading/unloading; “demurrage” is the terminal or storage fee charged when a container overstays its free time. ↩︎
-
For an example of an algorithmic strategy leveraging the exploration vs. exploitation trade‑off from an economic perspective, see “Multi‑armed Bandit Algorithms and Empirical Evaluation” (2005) by Joannes Vermorel and Mehryar Mohri. ↩︎
-
In supply chain settings, raw materials, work in progress, and finished goods held in stock frequently dominate short- and midterm variations in working capital. The higher the working capital, the more financing the company requires to operate. ↩︎
-
A second degree of indirection is now emerging as large language models (LLMs) assist in writing and refactoring those recipes under the supervision of human experts. The responsibility does not move; only the typing accelerates. ↩︎
-
User Interface ↩︎
-
Continuous Integration, Continuous Delivery ↩︎
-
This methodology progressively emerged at Lokad in the early 2010s. Joannes Vermorel coined the term “experimental optimization” in 2021 to formalize the approach. ↩︎
-
“Strange women lying in ponds distributing swords is no basis for a system of government.” comes to mind as a famous line from the British comedy film “Monty Python and the Holy Grail” (1975). ↩︎
-
Recent progress indicates that much of the clerical work behind Fermi-style decompositions can be delegated to large language models operating in “deep research” mode: the model proposes factorizations, retrieves order-of-magnitude anchors with citations, and drafts calibrated ranges for human vetting. Benchmarks centered on decision-oriented forecasting—e.g., ForecastBench (2024), a continuously updated evaluation of models on probability-bearing questions—report competitive performance by modern models; in parallel, methodological studies show that careful prompting materially improves calibration and prediction intervals. Economics remains the binding constraint: producing millions of such Fermi-ized ranges per month is not yet sensible for most firms, yet generating a few thousand well-documented ranges monthly is already operationally straightforward with today’s APIs and agentic tooling. ↩︎
-
“Fourth-generation languages” promised, from the late 1970s onward, to let analysts describe business logic declaratively—report writers and data-manipulation languages tethered to a database. The label faded by the early 2000s, but the underlying need did not. ↩︎
-
Visual Basic for Applications ↩︎
-
In computer programming jargon, a heisenbug is a software bug that seems to disappear or alter its behavior when one attempts to study it. The presence of heisenbugs usually reflects the lack of time-travel reproducibility. ↩︎
-
Edgar F. Codd introduced the relational model of data in 1970. His landmark paper “A Relational Model of Data for Large Shared Data Banks” remains the foundational reference for relational databases. ↩︎
-
Kenneth E. Iverson is credited with developing array-based approaches to programming, beginning with APL in the 1960s. ↩︎
-
APL, MATLAB, NumPy, etc. ↩︎
-
Early industrial uses of distributed data processing can be traced back to mainframe clustering in the 1980s, but frameworks like MapReduce (2004) and Apache Spark (2014) popularized more accessible paradigms for large-scale dataflow computations. ↩︎
-
REPL stands for Read-Eval-Print-Loop, a programming environment in which code can be typed, immediately executed, and the output returned in real time. ↩︎
-
See earlier discussion of how materials such as cables already defeat most off-the-shelf inventory tools, indicating how minor-looking quirks often break rigid analytical models. ↩︎
-
Companies frequently distinguish RFIs (Request For Information), RFPs (Request For Proposal) and RFQs (Request For Quotation). For concision here, “RFP” refers collectively to RFI, RFP, and RFQ—all three equally dysfunctional for the same reasons. ↩︎
-
As a litmus test, any consultant agreeing to support an RFP for enterprise software should be dismissed outright for dangerous incompetence. ↩︎
-
Loosely inspired by Warren Buffett’s “silver bullet” question: If you could get rid of one competitor, who would it be? ↩︎
-
Joel Spolsky, a famous American Software Entrepreneur, published an article “Good Software Takes Ten Years. Get Used To It.” (2001). While this very broad statement is debatable, one must recognize that, a quarter of a century later, it remains essentially true. ↩︎
-
Called the Joint Procedure Manual internally at Lokad. ↩︎
-
It’s always 2×2 matrices. ↩︎
-
See “Leadership styles”, page 149, Dynamic Supply Chains (2010), John Gattorna. ↩︎
-
Any CEO knows that half of his employees are doing nothing of value, but he doesn’t know which half. ↩︎
-
Independence is a simplifying assumption. Correlations—e.g., shared capacity shocks—lower the joint entropy relative to the sum of per‑SKU entropies; they do not overturn the order‑of‑magnitude gap shown here. ↩︎