A general structure of the Haze has already been suggested by the title of Part I: A Confluence of Factors. Both the temporal juxtaposition of events, which eventually, albeit at an unknown point in time, results in the arena for discovery, as well as the experimental and theoretical scientific details contribute in differing yet interdependent ways (Fleck 1980, p. 128), (Graßhoff 2008). There is no “grand design,” neither in the minds of those ultimately responsible for the decisions nor in the outcome of individual experiments. It is an adaptive process during which we are led one step at a time. In reference to the rise of the Medici family in fourteenth century Florence, John Padgett, et al. wrote, “heterogeneity of localized actions, networks, and identities explains both why aggregation is predictable only in hindsight and how political power is born (Padgett and Ansell 1993).” If we replace “political power” with “a scientific discovery” the statement remains just as applicable.

There are many examples in the literature in which a dynamic is described with stages of progress, including basic scientific research and industrial and commercial development, that lead to technological innovation (Bush 1945a), (Rogers 1995, pp. 131–137), (Branscomb and Auerswald 2002). One classic description is Vannevar Bush’s “pipeline,” introduced in 1945 as a post World War II effort by the U.S. federal government to prepare basic research results for use in private industry. Another concept envisions a “valley of death” dramatizing the demands put on entrepreneurs during the shift from invention to innovation. A similar metaphor describes a chaotic “Darwinian Sea” inhabiting the divide between the shores of science/technological and investor/industry which only the most well-suited ideas may cross (Auerswald and Branscomb 2003; Bonvillian 2014; Branscomb and Auerswald 2002; Bush 1945b). A more intricate example is “the chain-linked mode of innovation,” a reaction to theories describing the innovation process as having some kind of conceptual order in the hopes that clearer organization would aid in policy development. Here, multiple feedback loops exist between scientific research, technological innovation, product design and marketing as opposed to a linear process (Kline and Rosenberg 1986; Sanderson and Uzumeri 1995).

The interdependence of scientific (or basic) research and technological innovation is rightly emphasized in many of these models, especially the more recent ones (and has been a point of criticism toward Bush’s pipeline (Brooks 1994; Langrish 2017)). However, while some models of innovation view scientific research as the earliest possible step (one that is not always necessary or may be revisited through feedback loops), it is sometimes missing altogether, even in cases where it plays a role. Descriptions of the mechanism of scientific progress itself are even scarcer (Rogers 1995, p. 132). Rather, a successful scientific outcome is often used as a starting point, or at least presupposed. In itself, the point in time immediately following a scientific breakthrough is not an invalid place to begin a discussion of technological innovation, but a discussion of the preceding scientific progress would be beneficial because scientific discovery and technological change are inevitably intertwined. Unfortunately, attempts at an explanation can fall victim to the same oversimplification that innovation researchers themselves warn against when describing technological advances.

As was shown in the case of ammonia synthesis, normal scientific research is as continuous as innovation in technology and change in other fields. The notion is widespread but is not always applied to science (Schumpeter 1952, p. 89), (Globe et al. 1973), (Stigler 1982), (Becker 1984, pp. 301–310), (Chandrasekhar 1987, p. 14), (Basalla 1988, pp. 30–57). “The final product of innovative scientific activity,”Footnote 1 wrote George Basalla, “is most likely a written statement, the scientific paper, announcing an experimental finding or a new theoretical position (Basalla 1988, p. 30).” However, normal science thrives on the same additive dynamic as does technological innovation and often a breakthrough represents no more a conceptual leap than the results of an “ordinary” experiment. Only the context is different. A publication is no more the “end product” of science than the iPhone is the end product of cell phone research (although they may misleadingly appear to be more than they are (Rudwick 1985, pp. 434–435)). Fritz Haber’s 1905 publication on ammonia synthesis, if this were considered to be “the paper,” fits into a continuous scientific narrative on the development of classical thermodynamics and its transition to a basis of statistical and quantum mechanics. Energy science in the mid-1800s led to an understanding of heat and work which was applied to gases; this knowledge was adapted to chemical solutions and gave birth to physical chemistry. Subsequent steps came in the form of Walther Nernst’s third law with experimental confirmation by Fritz Haber. The closure of classical thermodynamics then provided an avenue for the application of more modern physical theories (statistical mechanics and quantum theory) to these same systems. Each of the achievements was marked by one or more publications but did not necessarily result in a precipitous jump in quantity or quality of physical knowledge. While each achievement resulted in improved abilities of prediction (and even clairvoyance), they also suffered from inaccuracy and confusion.

Scientific research has also been treated analogously to technological innovation in that different, smoothly varying outcomes are assumed possible (Rudwick 1985, p. 450). In what is, incidentally, a comparatively broad discussion of scientific progress, Kline and Rosenberg discussed Galileo’s contribution to mechanics (Kline and Rosenberg 1986). “Without the telescope,” they wrote, “we would not have the work of Galileo, and without that work we would not have modern astronomy and cosmology […] It is probable also that without Galileo’s work we would not have had what we now call elementary mechanics until a much later date, and perhaps not at all [italics added].”

Here again, the idea of the heroic inventor in technological innovation is invoked, or the heroic scientist, the general centrality or inextricability of whom I have argued against because it is so incompatible with continuous progress. While they do appear, sometimes to great effect, the success of a single person is often due to (ideal or less than ideal) circumstances outside their control or they are later made into something they never were (Fleck 1980, p. 61), (Holdermann 1960, pp. 95, 96), (Kasperson 1978; Simonton 1984), (Becker 1984, pp. 300–301), (Rudwick 1985, p. xxii, 15, chapter 2, pp. 411–428), (Basalla 1988, pp. 57–66), (Padgett and Ansell 1993), (Csikszentmihályi 1996, pp. 330–336), (Adichie 2009). One reason given is nationalism, which is certainly true in the case of Fritz Haber. Another is the confusion when continuous technological change leads to outsized social or economic consequences (the “catalytic effect”) (Rudwick 1985, p. 438), (Barley 1990). Again, this is true with Haber. His scientific contribution was not outsized and, in fact, Nernst’s theoretical solution held greater scientific consequence. But Haber’s was the final step before the power of physical chemistry could be unleashed in the form of the high pressure catalytic chemical industry. The circumstances shaping the arena for discovery were decisive.

Returning to Galileo, did Kline mean to say it is probable that if it were not for Galileo, we would now have an entirely new theory of mechanics? I think we would not. There is again the argument against the hero. “Galileo gave to modern science the quantitative experimental method,” wrote Hans Reichenbach. “Yet this general turn toward experimental method can scarcely be regarded as the effect of one man’s work. It is better explained as the result of a change in social conditions…[that] led naturally to an empirical science (Reichenbach 1954, pp. 89–99).” It is certainly possible that without Galileo the development of mechanics would have been set back, or, with a substantially smaller probability, that we would have no theory at all. Where would that leave us after having entered a planetary reality of “an infinite number of mutually separated and isolated, hard and unchangeable—but not identical—particles (Koyré 1968)”? Truly with no theory of mechanics for all times? Would we eventually develop an altogether new theory? We currently have three options to choose from: “classical” Newtonian mechanics, quantum mechanics, and relativistic mechanics. While it is clear to us today that the first is an approximation that works at the dimensions and speeds of the reality occupied by the human being, it is a very good approximation and is substantiated by the other two theories if appropriate constraints are applied. We still rely on classical mechanics for many calculations—at our order of magnitude there is still no better or simpler option. And no one expects to find one. Considering our place in the universe, it is unlikely quantum mechanics or relativistic mechanics would have been developed before classical mechanics as the quantum of action and the invariability of the speed of light are needed for these theories. Likewise, it is improbable that observations made by human-like beings working under the assumption that the universe exists in a state similar to the one with which they perceive to directly interact would have led them to a description wholly different than classical mechanics. Velocity or momentum may have been used as basic units instead of distance and time, but the underlying physics of the interaction of solid objects moving in the way we humans can most easily observe and contemplate would have remained the same. Having said this, if we were the size of an electronFootnote 2 we may first have had a functioning quantum mechanics (which we would know simply as “mechanics”) while current scientists may be slogging away on a theory of “continuous mechanics” to describe the mysterious world of huge conglomerations of matter on the unfathomable order of magnitude of kilograms. Would our observations have already reached the scale on which relativistic mechanics is needed? We find ourselves at dimensions between the domains of the quantum and the astronomical. Both realms have become accessible to us, making it a unique vantage point. Or is it? Perhaps it is a normal step beyond the scales that one inhabits to understand how to observe the scales above and below and we are still not able to see past the initial step in each direction. The paradigm we use depends not only on how we choose to see the universe, but how we are able to see the universe; the theory or paradigm that results from research is dependent on the way the universe (or system) appears to be as it is developed. Without the work of Galileo himself, therefore, I think we would still be in control of a classical theory of mechanics because of the kind of instinctively “natural” universe it describes to us. In contrast, quantum mechanics is almost a century old and it has still not been grasped by most physicists. Could such an absurd view of the universe really precede classical mechanics?

If we were able to peer further or observe more, we might discover a form of contemplation different from anything we already know, but for each perspective there is only one paradigm that fits. By interpreting an observation in a certain way, a useful theory may be built upon it, though a different interpretation may show that theory to be limited, as surely all theories are (or will eventually be shown to be) with respect to some perspective on the natural world. No one contends that any single paradigm they employ describes all aspects of the universe in a perfect manner. But a useful paradigm contains representations of the truth and a set of rules that allows for communication and continuity in research, even after a new paradigm achieves greater precision or more accurate descriptions. “Ptolemy’s astronomical system,” wrote Reichenbach, “also called the geocentric system, is still used today to answer all those astronomical questions which refer merely to the aspect of the stars as seen from the earth, in particular, questions of navigation. This practical applicability shows that there was a large measure of truth in Ptolemy’s system.” The only drawback was “the imperfect state of the science of mechanics at the time (Reichenbach 1954, pp. 97–98).” Even though we now have a more mature theory of mechanics, it is conceivable to generate new knowledge within Ptolemy’s outdated paradigm and create new technology, perhaps a consumer navigation device for the hobby sailor. Physical chemistry, as Haber employed it, was also very useful, even in its classical form; later it was extended to reveal more about the behavior of nature. Its first incarnation is, however, not illegitimate.

The kind of paradigm or theory which functions correctly (however naive or incomplete it may be) must be based in some way on the “spirit” of science. There must be some observation and an effort at interpretation using physically insightful and meaningful tools that deepen the meaning of the observation. That the sun rises and travels across the sky is a meaningful observation, for example. That the greek god Helios drives his chariot each day toward Oceanus is not a meaningful interpretation; it does nothing to explain or embed itself in the further workings of the universe, except, of course, that our world is inhabited and manipulated by various gods, specialized in their own self-absorbed rituals. This “fact,” however, will never be substantiated by any observation because it is not true. There are also aspects of the scientific perspective of the universe, including energy conservation and the classical force of gravitation, that cannot be derived from more integral, basic ideas. However, they are based on repeated observation, give us insight into how the universe may function or behave, and allow us to generalize our observations to derive further useful knowledge (while keeping in mind that these are stated “laws” and not proven fact). As long as a theory operates in this way, it can be considered science and one may say that it contains a “measure of truth.” That is why, scientifically, Helios is no longer of any use to us.

Later, the role of different paradigms will be considered further, but first we return to where we began: normal science within a paradigm, which, in contrast to technology, is adorned with objectivity. The application of electromagnetism to the electric motor, for example, did not predetermine its final configuration (Basalla 1988, p. 43). The view of electromagnetism of mutually-inducing electric and magnetic fields did, however, lock in Maxwell’s equations. Every subsequent derivative description followed from there.

When considering science, the objectivity within a given framework is a commonly misunderstood characteristic. Science is often viewed by the general public as thoroughly objective and dealing in hard facts; within other expert or academic circles, the same scientific activity is often considered, at some level, to share the subjectivity found in technological innovation or progress and change in medicine, politics, art, sports, business, and other fields. It is neither of these. Normal science is subject to strict rules. The way in which the rules must be followed, however, is limited only by creativity and ingenuity. Like many endeavors, it is an art form. The goal of the scientist, be it the creation of knowledge for the common good, the pleasurable or perhaps selfless “need to know,” or the achievement of lasting acclaim, may seem only fleetingly definable. “I admit,” wrote Subrahmanyan Chandrasekhar, in trying to capture the essence of science, “that these are things which cannot be defined any more than beauty in art…(Chandrasekhar 1987, p. 13 and chapter 4).” Add to this nebulousness the challenges of developing a new physical theory and the complexities of science seem to escape the simple two-part description consisting of normal science and paradigm shifts.

In describing the Haze further, I focus on three factors that illustrate science and scientific discovery in a way that reveals the overlap and interdependence of fact-gathering and paradigm shifts and how a scientific breakthrough is the result of this relationship. In doing so, I maintain the definition of “science” as the interpretation of experimental results, obtained in a transparent and reproducible way,Footnote 3 within a theoretical framework. The three factors are:

  1. (1)

    The existence of (what behaves like) an objective truth—a fact or optimized set of system conditions—within a given paradigm. This objectivity results from the rigorously defined way a particular aspect of natureFootnote 4 is perceived and treated by researchers bound to that paradigm. The word “subjective” refers to activity not fully within this context.

  2. (2)

    The environment of knowledge exchange within which a scientist works includes living colleagues but also the paradigm—a body of knowledge built up over generations—which provides a framework in which they act. Whether a scientist is interacting with his colleagues (social interactions) or facilitating an exchange between established theory and experiment (epistemic exchange), the successful transfer of scientific knowledge depends on the objectivity in 1).Footnote 5

  3. (3)

    The value of scientific results, concepts, equations or laboratory procedures is not immediately or completely identifiable, nor is there a general chain of development by which value is increased. Rather inherent value is recognized or assessed.

I will not attempt to elaborate on the value of science here, but a few words will help explain my use of the term.

Scientific value derives from an advance that allows or forces us to view a system in a new or more fundamental way (Siggelkow 2007)—that is, it makes us think differently (Fortus 2018). The only immediate limitation is our creativity when combining theory and experiment. The inability to think outside the box has proven to be a real problem. As Gerhardt Ertl put it while contemplating the cleverness of the van’t Hoff Box in Appendix A (Ertl 2018): “We have the money—what we need now are ideas!”

Polanyi identified three contributions to value in science: certainty (accuracy), systematic relevance (profundity), and intrinsic interest (Polanyi 1962, pp. 135–136). The first two can be grouped together into what I call “usefulness” with identifiable value within a paradigm. The third, a subjective factor, plays a role in the identification of pertinent scientific questions along with “our vision of reality, to which our sense of scientific beauty responds.” However, while intrinsic interest may make something fascinating, it does not make it scientifically valuable.

I find the best way to express the value in science is as the integral of all usefulness over time (which may be positive, negative, or zero):

$$\displaystyle \begin{aligned} Scientific\ Value = \int_{t=\tau_o}^{Present} Usefulness \cdot dt {} \end{aligned}$$

All that remains is to define the time-dependent function “Usefulness.″ It will not be a straightforward task; we may just have to get creative. One widely considered possibility that must be discounted stems from the detrimental misconception that productivity and proliferation alone correlate with usefulness and scientific value. They do not—except perhaps inversely.

Returning to the three factors above, I again stress they are not meant to suggest the primacy of a scientific discovery over a technological innovation. While there are many innovations which have occurred seemingly without direct input from basic scientific research, it would be erroneous to declare any modern scientific discovery came about without the aid of technology. I am also not suggesting the superiority of the scientific method over approaches that have been successfully employed in other fields. Rather, I am stating that the scientific method has enabled significant progress in some areas. Here, the focus is on the type of scientific discovery that could be viewed as a decisive event or “an especially important event that provides a major and essential impetus to the innovation. It often occurs at the convergence of several streams of activity. In judging an event to be decisive, one should be convinced that, without it, the innovation would not have occurred or would have been seriously delayed (Globe et al. 1973, p. 2).” Such a decisive event need not be a scientific discovery, but it certainly may be. Conversely, a significant scientific discovery need not lead to technological innovation. This situation, in the context of ammonia synthesis, represents the most inclusive process chain leading to innovation, though the presence or absence of technological consequences has no effect on the dynamic of the Haze itself. Prior to the arrival of the decisive event (in this case, the scientific breakthrough) there are a series of occurrences which, to varying degrees, may be drawn into a relationship to the discovery. These may be classified as nonmission-oriented research (NMOR), or “research carried on for the purpose of acquiring new knowledge, according to the conceptual structure of the subject or the interests of the scientist, without concern for…application” and mission-oriented research MOR, or “research carried on for the purpose of acquiring new knowledge expected to be useful in some application (Globe et al. 1973, p. 3).” It is this change from overwhelmingly NMOR to MOR with respect to a specific breakthrough with is embodied by the confluence of factors. It is this dynamic which illustrates the condensation of the Haze into the arena for discovery and the precipitation of the breakthrough.Footnote 6 It is this arena that separates the preceding “normal” fact-gathering from a subsequent scientific discovery. The individual steps of scientific advancement may appear identical in both cases with only the context supplying a means of differentiation. It is not a linear convergence because feedback mechanisms interlink aspects of basic science and technological progress. During the condensation, uncertainty and differences of opinion among the practitioners of a discipline concerning the exact knowledge defining a breakthrough are vastly reduced, the spread of choices narrowed (Fleck 1980, pp. 15–16, 72–73, 110), (Rogers 1995, pp. 13, 36–37), (Rudwick 1985, p. 455), (Kline and Rosenberg 1986; Padgett and Ansell 1993). Available information allows the astute observer to identify which experiments (and which aspects of those experiments) are necessary to articulate the paradigm in the way that reveals the breakthrough, as if bringing closure to the acute questions that first exposed the arena for discovery. Let us consider the events leading to ammonia synthesis from the elements in Fig. 17.1. There are certainly external factors that shape the trajectory of a discovery, such as the Margulies brothers offering Haber financial compensation for his research or the academic political shuffle within which Haber and Nernst were embroiled. Here, however, we restrict ourselves to scientific developments.

Fig. 17.1
figure 1

The scientific developments during the confluence of factors leading to ammonia synthesis and their relationship and relevance to work between 1903 and 1908 by Fritz Haber, Walther Nernst, and others. NMOR stands for nonmission-oriented research, MOR for mission-oriented research. See text for details

The transition from NMOR to MOR is particularly clear in this case, as is the convergent nature of the NMOR or MOR on its own. Consider, for example, Haber’s initial results in 1905 compared to the increased accuracy that followed (as well as his reinterpretation of the initial results in terms of the increased knowledge base described in Part II, Chap. 11). The accumulation of MORs near the discovery—not necessarily in a temporal, but rather in a knowledge-based sense—may be a consequence of the difficulty of actually achieving the breakthrough, in other words, of properly embedding the last step in an attempted paradigm articulation within the complex network of knowledge that preceded and led to it (Klein 2016a). “It happens again and again,” wrote Wilhelm Ostwald in 1908 (Ostwald 1908, pp. 22–24),

…that of all things, the very last step, with which the new thought would be completely perfected and able to stand in opposition to old ways of thinking, is usually forgotten, overlooked, or disregarded by the creator of the new idea.Footnote 7

While Ostwald may have been referring to groundbreaking “new thought” more in line with paradigm change, the travails of normal science will not have been far from his mind: just a few years earlier, he had concluded his failed attempt at ammonia synthesis. After Carl Bosch reported his inability to reproduce Ostwald’s experiments at BASF in 1900 (Part II, Chap. 9), the latter was “so exhausted that I could no longer bear further engagement with these things (Ostwald 1903, p. 287).”Footnote 8 Continuing with this line of thought, again in 1908, Ostwald commented that the aforementioned “creator” had “no more remaining strength” to finish his task.Footnote 9 He had been close to his goal and had defined the steps Haber would later successfully implement. It was only the attention to one detail, that minuscule and reproducible quantities of ammonia should result from his experiments, that evaded Ostwald’s consideration. Had he recognized this experimental possibility, he would have fully embedded his results within the framework of physical chemistry. In fact, the other researchers investigating ammonia synthesis at the end of the nineteenth century suffered similar fates. Ramsey and Young also measured values that were too large, whereas Le Chatelier lacked the necessary experimental ability when working with high pressures. Edgar Perman was an exception; in disregarding the principles of physical chemistry, he tried to articulate the wrong paradigm altogether. All of them were working in a time when the complete set of puzzle pieces was available. They simply had to be assembled in the right way—a task that can be as formidable as identifying the pieces in the first place. They must be assembled perfectly, of course, because there is no workaround. The objectivity of science places strict constraints on the acceptability of the final answer.

It is certainly not a general property of the Haze that it exhibit this level of simplicity, which again shows why ammonia synthesis is such an informative example of scientific discovery. In a general model of scientific progress, there must be the allowance for greater oscillation between NMOR and MOR and less demarcation of discrete, independent classifications of science so that any generic form of the Haze remains arbitrary. Certainly traditional, cultural, societal, political, or any number of external factors (including luck) also influence the speed at which the Haze condenses. They do not, however, have any influence on what precipitates.Footnote 10

During the transition period, there is a recognition of scientific value when the NMOR finally gives way to MOR and results become consciously applicable to further research. At this point, the position and relevance of the knowledge within the paradigm becomes apparent as does its usefulness, although a specific determination of value is only in regards to the corresponding discovery. Any NMOR, at an earlier date, may have given way to MOR (with respect to another subject) and any new MOR (with respect to ammonia synthesis) may already be considered MOR with respect to another application. Again, we are confronted by the delineation of events and can see why a schematic approach to one discovery helps isolate the underlying dynamic. This approach of assessing the value of science is abstracter than outcome-based methods (for example, assigning greater value to research that results in patents), but it is more fitting with respect to basic research (Rogers 1995, p. 135).

The procession toward scientific discovery can begin without any “need” for a solution. Later, after the MOR has evolved out of the NMOR and has made a principle contribution to the creation of the arena for discovery and proper insight has led to the breakthrough, revision is no longer required. Rather the revision comes just before the breakthrough as the MOR “circles” the correct answer and the scientists are aided not just by knowledge and expertise but also by intuition. This development is unlike technological innovation, where the step from invention to innovation is marked by the identification of a need and often requires a process of revision before the resulting product is market-ready (Basalla 1988, p. 23). Interestingly, the reason why a new technology (invention) may come to be seen as an innovation can be due to the existence of the technology itself (Langrish 2017; Sgourev 2015). That is, the perceived need for an innovation can either precede or follow the invention leading to it. In a scientific breakthrough, the arena for discovery where a distinct and definite problem has been revealed (to which a solution surely exists), having followed from the confluence of factors, must precede the successful breakthrough. The same is true even if the discovery is not recognized at the time the original work is completed; in such a scenario, the arena is not yet mature and the full articulation of the paradigm and resulting realization that a breakthrough has occurred is not yet possible. The arena does not identify need or market potential, it is the assemblage of knowledge and experience that reveals a relationship in nature according to the paradigm. From a scientific perspective, one particular result is not “needed” more than any other. The last puzzle piece is often crucial because it is the last, not because it is the most important. In this way, scientific results are not the same as technology—if a scientific result has value, it will eventually be identified through the mere circumstance that we continue to perform scientific work. Its place within the existing knowledge will be found. This activity inherently pushes us toward appropriate questions and, given enough time, the answers to those questions (Polanyi 1962, p. 135). It is as if something were nudging our understanding in a direction that is clear only in hindsight while the answer is still hidden in the paradigm. Can we represent this vague notion more concretely?

Our path through the research on ammonia synthesis would have gained considerable complexity with increased reference to personal relationships or local and international politics; this information would have added further insight and contextualization to the science, but it would have detracted from the opportunity to exploit the “relative simplicity” of the discovery to provide a simplified view of the otherwise complex structure of the Haze. Focusing on the science allows us, for example, to derive the schematic diagram in Fig. 17.2.

Fig. 17.2
figure 2

Schematic diagram of The Haze. The condensing cone on the left represents the confluence of factors leading to an arena for discovery in which the breakthrough, a unique and irregular structure, is precipitated. These together comprise the Haze. The right-hand expanding cone represents the consequences of the discovery, in particular the technological change that becomes ever-more interlinked with elements of our lives. The inset shows a close up of the breakthrough and its intricate structure. Image by Birgit Deckers at the Max Planck Institute for Chemical Energy Conversion

The condensation of the Haze may be represented with a cone whose axis lies parallel to the direction of time, growing denser as it narrows. The point of the cone marks the beginning of the arena for discovery in which the breakthrough is precipitated. In general, it is an episode drawn out in time (not a single event) with a unique, irregular structure as illustrated by the events of Part II. If the scientific breakthrough is a decisive event for technological innovation, as it was with ammonia, the subsequent developments may be represented with a second cone also lying parallel to the time axis; its point is at the end of the breakthrough and extends away from this moment, becoming dilute. The expansion contains technological developments that permeate and reconfigure our daily lives (Padgett and McLean 2006). It represents a broadening effect over time, not necessarily a “catalytic,” non-linear reaction, and may include desirable or undesirable, direct or indirect, as well as anticipated or unanticipated consequences (Rogers 1995, pp. 30–31), (Sveiby 2017). In the case of ammonia synthesis, some of these are the production of industrial fertilizer, explosives, pesticides, and poison gas, as well as the high pressure catalytic industry. The expanding cone also contains further scientific progress. New ideas amend or replace older ones, such as the application of statistical mechanics or quantum theory to thermodynamics (Kuhn 1970, p. 21), (Fleck 1980, p. 29). This type of expansion of a novel concept or device is not limited to science; it is seen across the spectrum of human activities (Brooks 1973; Bush 1945a; Sgourev 2013, 2015).

The use of a cone is indeed a stark simplification of the actual structure of the Haze, but its solidity reminds us that holes or jumps in advancement are rare; the fluidity is a consequence of the continuous nature of scientific and technological advancement. The difficulty is, rather, the way in which we choose to isolate events that are at once connected but also in some way clearly separate. It is a problem known in defining different classes of technologies or, more generally, different artifacts of many kinds (Murmann and Frenken 2006; Vedres and Stark 2010).

The double-cone structure in Fig. 17.2 may appear as a linear approach to a non-linear dynamic but only because the complexity has been hidden away for ease of management. The trick of simplicity to achieve transparency is an old one:

Every concrete process of development finally rests upon preceding developments. But in order to see the essence of the thing clearly, we shall abstract from this and allow the development to arise out of a position without development (Schumpeter 2012, p. 64).Footnote 11

Put another way: “the main direction of the development, taken as an idealized average, would have to be drawn separately and at the same time [as all other lines of development] (Fleck 1979, p. 15).”Footnote 12

An objection to this schematic diagram is the neglect of influence of prior or external science and technology on the condensation of the Haze. To represent this interaction we can draw as many overlapping structures as needed in an ever-widening array (Fig. 17.3). Some may be double cones while others are single structures consisting of only the Haze (a scientific discovery not immediately leading to a new technology), or an expanding cone (a new technology not dependent on a scientific discovery). A complete picture would consist of an unwieldy number of cone structures extending the length of any relevant time axis. This notion would not help in understanding the underlying nature of a single breakthrough, but it elucidates the sophisticated interaction between science and technology as well as the problem of delineation: a stand-alone object that also exhibits a strong dependence on scientific, technological, and other events.

Fig. 17.3
figure 3

Illustration of the interdependent nature of scientific discovery and technological innovation based on the schematic diagram of the Haze in Fig. 17.2. Three structures are present: the double-cone, the condensing cone (the Haze) representing a scientific breakthrough not immediately leading to a new technology, and the expanding cone representing a new technology not dependent on a scientific breakthrough. In reality, the density of shapes is much greater

Later, one addition to the schematic double-cone diagram will be discussed: a transitional stage between the breakthrough and the expanding second cone, representing hybridized scientific and technological work (Fig. 19.1).