National Academies Press: OpenBook

New Directions for Chemical Engineering (2022)

Chapter: 2 Chemical Engineering Today

« Previous: 1 Introduction
Suggested Citation:"2 Chemical Engineering Today." National Academies of Sciences, Engineering, and Medicine. 2022. New Directions for Chemical Engineering. Washington, DC: The National Academies Press. doi: 10.17226/26342.
×

2
Chemical Engineering Today

The work of chemical engineers has transformed societies and individual lives around the world, particularly in the United States. Chemical transformations are at the heart of the technologies that enable modern society (Box 2-1). Without synthetic fertilizers made with chemical engineering processes, Norman Borlaug could not have led a Green Revolution to feed the world. Without the invention of catalysts due to the leadership and vision of Karl Ziegler, a chemist, and Giulio Natta, the first chemical engineer to win a Nobel Prize, polyolefins would not exist, and the myriad benefits of plastics would not have been realized. Without the invention of tough, stable polymers by chemical engineers at the DuPont Company, including Roy Plunkett and Stephanie Kwolek, Teflon® and Kevlar® would not have been developed, nor would any of the commercial and medical devices made from those polymers. Without the contributions of chemical engineers such as Andy Grove at Intel and numerous others, the silicon chips, glass materials, and plastics that make up today’s ubiquitous electronic devices would not have been created. Without the leadership and vision of Nobel Prize winner Frances Arnold, who built upon the discoveries of numerous biologists, biochemists, and chemical engineers, the tools of directed evolution would not have emerged. Without an army of chemical engineers, there would be no oil and gas industry to power the world. And without chemical engineers, a robust pharmaceutical industry would not have been able to discover and produce the therapies and vaccines needed for a long and healthy life. Figure 2-1 highlights some of the areas in which chemical engineers have made major contributions.

At the same time, however, chemical engineering is also responsible for unintended consequences, such as those resulting from the production of chemicals that will persist in the environment indefinitely, greenhouse gas emissions that contribute to climate change, plastic materials that accumulate in landfills and the oceans, and the chemicals of war that have inflicted long-term or permanent damage on humans and the environment. Thus the field of chemical engineering today faces challenges and opportunities not only to innovate for the future, but also to innovate in ways that repair the unintended consequences of the past.

Suggested Citation:"2 Chemical Engineering Today." National Academies of Sciences, Engineering, and Medicine. 2022. New Directions for Chemical Engineering. Washington, DC: The National Academies Press. doi: 10.17226/26342.
×
Suggested Citation:"2 Chemical Engineering Today." National Academies of Sciences, Engineering, and Medicine. 2022. New Directions for Chemical Engineering. Washington, DC: The National Academies Press. doi: 10.17226/26342.
×
Image
FIGURE 2-1 Schematic illustrating the broad impact of the chemical sector on several aspects of American life. Examples of basic chemicals, specialty chemicals, agricultural chemicals, pharmaceuticals, and consumer products are shown. SOURCE: DHS (2019).

Chemical engineering as a discipline brings together the three fundamental sciences of chemistry, physics, and biology, as well as mathematics. Chemical engineers are agnostic to the material used or the particular application; they work with all phases of matter—vapor, liquid, supercritical, solid, and plasma—and with multiphase mixtures and at interfaces. They work at and across all length scales, ranging from molecules to medicines to materials and even machines. Their work creatively transforms matter and products into higher-value materials and products using the principles and tools of thermodynamics, transport, kinetics, process control, and process design. Chemical engineering is the only field of engineering that takes advantage of chemical transformations, usually followed by separation and purification, to add value to products.

Chemical reactions and reacting systems are central to many transformations. The problems tackled by chemical engineers usually involve time dependence, transport phenomena, and nonequilibrium phenomena and feature many variables. Chemical engineers are trained to take a system-wide approach to solving problems, recognizing that a single step is rarely adequate to complete the desired transformation. They realize that a system of connected components needs to be optimized holistically to be most efficient and economic. Importantly, chemical engineers appreciate the balance among physical performance, fiscal profitability, and safety.

The education and training unique to chemical engineering produce intellectually versatile engineers who are comfortable with mathematics and computers, skilled in analytical thinking, and adept at solving open-ended problems. They draw on an education that has served them well in diverse areas of accomplishment. Accordingly, chemical engineers are found in interdisciplinary teams in the agricultural, biological, biomedical, chemical, environmental, food, health, energy, materials, petrochemical, nanotechnology, pharmaceutical, and semiconductor industries, as well as in the consulting, communications, software, financial, and insurance sectors. They engage at all levels in these indus-

Suggested Citation:"2 Chemical Engineering Today." National Academies of Sciences, Engineering, and Medicine. 2022. New Directions for Chemical Engineering. Washington, DC: The National Academies Press. doi: 10.17226/26342.
×

tries and sectors, from research and development, to analysis, administration, and leadership, and they contribute to the engineering enterprise from scientific discovery, to application to scale-up, to commercial deployment, to plant operations.

THE DISCIPLINE

Chemical engineering is both a discipline and a profession. Its academic disciplinary roots trace back to the 1880s. In 1880, the English entrepreneur George Davis proposed the creation of a Society for Chemical Engineers, which resulted in the creation of the Society of Chemical Industry; in 1886 and 1887, diploma curricula began at City and Guilds College (now Imperial College) and at Glasgow and West of Scotland Technical College, respectively; and in 1888, Lewis Mills Norton proposed the first course in chemical engineering at the Massachusetts Institute of Technology (MIT, 2021). Chemical engineering evolved as a profession from the roots of industrial and applied chemistry, which in turn emerged from such ancient chemical processes as fermentation and leather tanning. Traditionally, industrial and economic needs and academic research have been tightly coupled, with many faculty members being engaged in consulting or joint research work with industry. To paraphrase L. E. Scriven, the academic tree takes its nourishment from the soil of practical challenges and periodically returns its many leaves of results and trained graduates to that soil (University of Minnesota, 2003).

This discipline is the only one in engineering with a central focus on molecules and their transformations—that is, chemistry. For much of its history, chemistry and physics were the core sciences of chemical engineering (along with mathematics). Because of its fundamental engagement with chemistry, it is natural for chemical engineering to be the principal home for the engineering of biological systems of all scales. The connection between chemical engineering and biology has a long history, and their interactions grew exponentially with the advent of methods for genetic engineering and the advancement of molecular biology from the late 20th century onward (Box 2-2). Indeed, numerous fields (e.g., biochemical engineering, metabolic engineering, tissue engineering, synthetic biology) have developed as outgrowths of chemical engineering or with key contributions from chemical engineers. The expansion of the influence of biology on the field has also been so significant that many academic departments have changed their names to incorporate some version of “bio” in addition to “chemical” engineering.

Because chemical engineers deal with both molecules and the enormous industrial plants that produce them, their work encompasses a large range of length and time scales, from the nanometer scales of chemical bonds and reactions to the kilometer scales of crude oil (petroleum) refineries, and from nanosecond chemical reactions to batch processes that take hours. Few fields of science or technology deal with changes of more than a dozen orders of magnitude in both length and time scales. The core curriculum of chemical engineering has lasted more than a century because it focuses on the analysis of linked processes regardless of scale.

The early history of chemical engineering has been well described (Colton, 1991; Furter, 1983). The growth of the core curriculum is marked by several events, the first of

Suggested Citation:"2 Chemical Engineering Today." National Academies of Sciences, Engineering, and Medicine. 2022. New Directions for Chemical Engineering. Washington, DC: The National Academies Press. doi: 10.17226/26342.
×

which featured George E. Davis in England in the 1880s. Davis was an entrepreneur in chemical manufacturing who called for the creation of “chemical engineering” to address the rampant pollution generated by chemical processes (Cohen, 1996). The first course (Course X) developed by Lewis Mills Norton at MIT had its stops and starts, but by 1902, William Walker had accepted the leadership of Course X while remaining in consulting partnership with Arthur D. Little until 1905. Arthur D. Little, as chair of the visiting committee for chemical engineering at MIT, is credited with creating the concept of “unit operations” in 1915 to describe the basic physical (and later, chemical) operations used to transform raw materials into products (Flavell-While, 2011).

Throughout the 20th century, further academic advances in the curriculum took place across the United States. At the University of Wisconsin, Olaf A. Hougen and Kenneth M. Watson wrote the three-volume Chemical Process Principles: Material and Energy Balances (1943); Thermodynamics (1947b), and Kinetics and Catalysis (1947a). Also at the University of Wisconsin, R. Byron Bird, Warren E. Stewart, and Edwin N. Lightfoot published the paradigm-shifting textbook Transport Phenomena in 1960. That book transformed chemical engineering by bringing a strong mathematical approach to the unification of treatments of fluid mechanics and heat and mass transfer, both reinforcing and explaining the connections made by Allan P. Colburn and Thomas H. Chilton at

Suggested Citation:"2 Chemical Engineering Today." National Academies of Sciences, Engineering, and Medicine. 2022. New Directions for Chemical Engineering. Washington, DC: The National Academies Press. doi: 10.17226/26342.
×

the University of Delaware and the DuPont Company, respectively, in 1946 (Chilton and Colburn J-factor analogy). At the University of Minnesota, Neal Amundson and Rutherford Aris worked to develop the analytic and mathematical foundations for reaction engineering. At the University of California, Berkeley, John M. Prausnitz developed methods of molecular thermodynamics beginning in the 1950s and 1960s. Throughout and after this period, the role of computing in chemical engineering control and design grew steadily (Box 2-3). Thus by the mid-1960s, a canonical undergraduate program, rooted in a foundation of fundamentals from chemistry, physics, and mathematics, was formed in a way that is easily recognized today.

THE PROFESSION

Chemical engineering as an industrial profession has its roots in processes for materials extraction and transformation. Éleuthère Irénée du Pont de Nemours founded his eponymous company in 1802 to make gunpowder on the banks of the Brandywine

Suggested Citation:"2 Chemical Engineering Today." National Academies of Sciences, Engineering, and Medicine. 2022. New Directions for Chemical Engineering. Washington, DC: The National Academies Press. doi: 10.17226/26342.
×

River in Wilmington, Delaware, at a site with a dam, a millrace, and saltpeter at hand. John D. Rockefeller became the world’s first oil baron when he united several companies into the Standard Oil Company in 1870, exploiting wells in Ohio, Pennsylvania, and eventually elsewhere. Standard Oil quickly came to control about 90 percent of America’s refining capacity (see Ohio History Central, 2021). Herbert Henry Dow founded his company in 1897 to extract bromine from underground brine in Midland, Michigan. Over time, each of these companies came to appreciate and need chemical engineering expertise and scientific advances to grow and diversify.

By the 1970s, the leading chemical engineering research efforts in U.S. industry were at Esso (later Exxon) and Mobil laboratories, DuPont Central Research and its Experimental Station, and Dow Central Research. Space does not allow enumeration of all of the advances in materials and processes that emerged from those laboratories and development efforts, or other important advances in research and product development from the laboratories at 3M, Shell, Universal Oil Products (later Honeywell UOP), Aramco, Standard Oil Company, Bell Labs, and elsewhere, but they all contributed in a significant and largely defining way to the quality of life enjoyed in the United States. Parallel developments also played out in chemical and other companies focused on materials. In 1953, for example, Lexan® polycarbonate was invented at the General Electric (GE) Company. This invention led to the creation of the GE Plastics Business, which grew to be a global business based on major contributions of chemical engineers (Plastics Hall of Fame, 2021).

Chemical engineering has also played an important role in electronic materials, a role amplified by the nearly ubiquitous role of electronic devices in today’s society. The expansive proliferation and adoption of electronics globally has driven an explosion of uses for electronics well beyond the traditional uses of personal computers and cell phones. Today, the introduction of 5G, the internet of things (IoT), cloud computing, and autonomous driving is having a direct impact on how electronics are used. Looking forward, the further adoption of machine learning and artificial intelligence to expand computational capability and demands for electronic devices are at an inflection point. While these trends had been progressing for some time, they have been accelerated by the COVID-19 pandemic, which has made technical advancement more urgent than ever—similar to the way it was in early days of the internet and the introduction of personal computers.

TIMES OF CHANGE

The world has changed and continues to change at an increasingly rapid rate. Technology is transforming the way everything is done, disrupting established practices and organizations. The range of challenges facing engineers is evolving, and the challenges are becoming more difficult. Engineering is about solving problems, so it is natural to expect that as one problem is solved, another emerges or grows in importance. The goal of education in chemical engineering is to equip graduates with the intellectual tools needed to solve problems and the ability to adapt those tools as new problems emerge. Information used to be valuable in its own right, but the internet has made information

Suggested Citation:"2 Chemical Engineering Today." National Academies of Sciences, Engineering, and Medicine. 2022. New Directions for Chemical Engineering. Washington, DC: The National Academies Press. doi: 10.17226/26342.
×

essentially free. What is now more valuable than ever is the knowledge needed to curate data and synthesize new and novel solutions, and to do so more rapidly and at lower cost than competitors. In a real sense, an undergraduate chemical engineering degree is a platform upon which its holder can build, with a research degree, a professional degree, and/or a career filled with formal and informal learning.

There is of course nothing new about change. The last time the National Academies surveyed the challenges and opportunities for chemical engineering—in the 1980s (Frontiers in Chemical Engineering: Research Needs and Opportunities, better known as the “Amundson Report” [NRC, 1988])—the conclusions reached suggested an approach to developing new technologies and maintaining leadership in established ones. A Centennial Symposium of Chemical Engineering (Wei, 1991) laid out the challenges facing the field at that time, and included a rich discussion of the balance of revolution and evolution in chemical engineering. Comments at the time about the tension between teaching the established core of such topics as transport, thermodynamics, kinetics, and design and making room for education relevant to (then) new areas such as biotechnology and materials science resonate today in this report. Nonetheless, the rate of change driven by an interconnected world is without doubt faster today than it was yesterday, and will be even faster tomorrow.

As new advances combine to produce even newer advances, the pace accelerates. At the time of the Amundson Report in the 1980s, personal computing was just becoming common, under 1 million cell phones were in use in the United States (Tesar, 1996), the internet was used by a cognizant few, and Microsoft had just gone public. The top ten public companies in terms of revenue included Exxon (1, at $91 billion), Mobil (3), Texaco (4), DuPont (7), and Amoco (10) (Fortune, 1985). The undergraduate chemical engineering curriculum focused on the core fundamentals of thermodynamics, transport phenomena, reactors and kinetics, process control, and process design. In 2021, as this report was being written, the internets of information and things were affecting all aspects of life, 14 billion cell phones were in use worldwide (Statista, 2021), the human genome was known and could be edited, it was possible to “see” individual atoms, and the use of artificial intelligence and deep machine learning was growing rapidly. The top ten U.S. public companies in terms of revenue in 2020 were Walmart (1, $514 billion), ExxonMobil (2), Apple (3), Berkshire Hathaway (4), Amazon (5), United Health Group (6), McKesson (7), CVS Health (8), AT&T (9), and AmerisourceBergen (10). After ExxonMobil, the next largest oil and gas company was Chevron, at number 15. Dow Chemicals, the largest materials science and chemical company in the United States, was 78th with revenues of $43 billion, less than 10 percent of Walmart’s (Fortune, 2020).

The undergraduate chemical engineering curriculum is still focused on the core fundamentals of thermodynamics, transport phenomena, reactors and kinetics, process control, and process design. The chemical industry, however, is a less important part of the overall economy than it once was, although the world can no more do without energy, materials, food, and water now than it could then. The top companies more than 100 years ago were all recognized industrial brands; today, technology companies have the largest market capitalizations. According to the Bureau of Labor Statistics (BLS), overall employment in chemical engineering dropped from 43,270 in 1997 to 30,120 in 2019, from

Suggested Citation:"2 Chemical Engineering Today." National Academies of Sciences, Engineering, and Medicine. 2022. New Directions for Chemical Engineering. Washington, DC: The National Academies Press. doi: 10.17226/26342.
×

almost 0.04 percent to shy of 0.02 percent of the national employment total (BLS, 2021). However, many people educated in chemical engineering work in areas not identified as such by BLS.

Over the last decade in the United States, the number of bachelor’s and master’s degrees awarded in chemical engineering more than doubled, a rate outpacing the 60–80 percent growth of engineering and STEM (science, technology, engineering, and mathematics) bachelor’s and master’s degrees. The number of doctorates awarded grew more modestly, with engineering growing most rapidly (Table 2-1).

Technology has also transformed the way people work. The linear industrial model of changing raw materials to products has shifted to a global interconnected platform model. That change, the increasing time-rate of change in society and business in general, and the substantial diminution of the economic significance of the chemical and oil and gas (but not the pharmaceutical) industries in the U.S. economy all have created substantial challenges and opportunities for chemical engineering.

The pace of change is now accelerating as a result of global connectivity, massive amounts of available data, artificial intelligence, sensors, robotics, and more. The background and training of chemical engineers are well suited to today’s rapidly changing world, and many of them have found their way into companies leading change. For example, as the new era of artificial intelligence and machine learning advances, chemical engineers are well equipped to move into these fields because of their strong background in reaction (network) engineering and control, along with an understanding of how to deal with complex interconnected processes, in addition to their fluency in math and computing.

These longer-term and probably irreversible societal and business changes are augmented by a critical need to mitigate climate change resulting from greenhouse gas emissions. The ultimate measure of success in addressing climate change will be reducing greenhouse gas concentrations in the atmosphere while still delivering the energy that society needs. The necessary system solutions align with the core chemical engineering practices of designing, building, and extending major manufacturing assets. Chemical process engineers have a long history of building safe, resource-efficient processes at the lowest capital and operating cost.

Along with addressing climate change and the energy transition are parallel needs—to reduce raw material usage and increase recycling to create a more circular economy; deal with the need to generate and distribute food worldwide while conserving water and other resources; and create and scale the manufacturing and distribution of new medicines and therapeutics. Each of these needs is discussed more fully in later chapters. There are other needs as well. Complicating efforts to address all of these needs are inevitable shorter-term geopolitical and environmental issues related to the price and availability of energy and feedstocks, as well as existing and emerging security threats and the viability of the installed asset base (Box 2-4). In these times of change, chemical engineers also play a central role in advances in many new and important areas, including personalized medicine, the infrastructure needed to produce vaccines for a pandemic, the rapid conversion to an economy based on nonfossil fuels, the need for new materials, and many more.

Suggested Citation:"2 Chemical Engineering Today." National Academies of Sciences, Engineering, and Medicine. 2022. New Directions for Chemical Engineering. Washington, DC: The National Academies Press. doi: 10.17226/26342.
×

TABLE 2-1 Bachelor’s, Master’s, and PhD Degrees Awarded in the United States between 2008 and 2019 in Chemical Engineering (ChE), All Engineering (Eng), and All STEM (Science, Technology, Engineering, and Mathematics) Fields

Field 2008 2009 2010 2011 2012 2013 2014 2015 2016 2017 2018 2019
Bachelors ChE 4919 5137 5838 6416 7176 7678 8202 9070 10032 11021 11653 11148
Eng 70232 70991 74778 78502 83636 88201 94386 100316 109373 118379 124794 114818
STEM 592,801 610,216 639,822 683,865 736,982 781,937 818,434 850,168 877,786 905,509 934,776 *
Masters ChE 937 996 1051 1284 1395 1453 1521 1629 1701 1801 1921 2003
Eng 33513 36909 38029 41751 43765 44037 46029 49855 55907 57754 56756 62682
STEM 164,007 175,895 184,097 199,386 215,137 225,261 235,368 249,762 274,084 290,037 298,157 *
PhD ChE 873 807 822 822 839 824 972 1002 920 930 981 1092
Eng 7860 7637 7575 8024 8452 8998 9585 9875 9455 9771 10179 12372
STEM 34,717 35,313 34,997 36,332 37,846 39,031 40,633 41,178 41,234 41,294 42,227 *

Data from ASEE (2020) and NCSES (2021).

Suggested Citation:"2 Chemical Engineering Today." National Academies of Sciences, Engineering, and Medicine. 2022. New Directions for Chemical Engineering. Washington, DC: The National Academies Press. doi: 10.17226/26342.
×

EDUCATIONAL CHALLENGES AND OPPORTUNITIES

While the core chemical engineering curriculum has apparently changed little over the preceding decades, the concepts are now taught in more modern ways. Initial unemployment for chemical engineering graduates is low and salaries remain high, although not at the highest levels relative to other fields, as was the case in the past. Thus, despite a course catalog that appears at first glance to be stuck in the past, the curriculum has been shown to yield graduates with the skills needed to adapt and succeed in the workplace. The hallmark of this continuing success appears to be the ability of chemical engineering graduates to think quantitatively, draw on data to guide the development of predictive models that can be expressed mathematically, and integrate pieces into a well-designed coherent system. In other words, “[the curriculum]…has endured not because it is frozen but because it has adapted dynamically to new ideas, emphases, challenges, and opportunities” (Luo et al., 2015). The undergraduate curriculum and chemical engineering education generally are discussed in more detail in Chapter 9.

New models of collaboration and integration are also emerging. Education has historically been designed as a linear flow from K-12 to college/university and then to a career. Today, global connectivity has led to new integrated learning and innovation practices. New education offerings are emerging from such online providers as EdX, Coursera, and Udacity. Online certificates and skills programs are finding wide application as the pace of change demands the constant refreshing of fundamental knowledge. The opportunity to build new shared content for use across universities is appealing, but would need careful consideration in light of the existing funding and reward systems within universities. Any such program needs to be available for lifelong acquisition and polishing of skills. Finally, assembling best-in-class content online for use as part of a university’s offerings can make content available to smaller or underresourced institutions, which in turn could open up new paths for engaging different and more diverse talent pools with the concepts of chemical engineering.

Suggested Citation:"2 Chemical Engineering Today." National Academies of Sciences, Engineering, and Medicine. 2022. New Directions for Chemical Engineering. Washington, DC: The National Academies Press. doi: 10.17226/26342.
×

For some, the undergraduate curriculum is a preamble to graduate studies. Graduate education is where the striking differences between the subjects undergraduates learn and the problems on which academics conduct research become clear. For example, undergraduates learn about small-molecule thermodynamics and some separation processes that have generally been highly specific to the oil and gas industry, although many bio-separation processes are now included in multiple courses. The oil and gas examples are obvious for historical reasons, but a vanishingly small amount of academic research is now occurring on those subjects. Thus, the real test for graduate students is their ability to adapt their critical thinking and analytic skills to problems in fields for which they probably lack both understanding of the vocabulary of the field and at least nontrivial basic science training.

This disconnect between the intellectual content of undergraduate and graduate work reflects the collective decisions of mainly federal but also other funding agencies that in large part support the work of graduate students. There is little federal support for work on basic thermodynamics or transport phenomena that an undergraduate would recognize. The disconnect is clear from even a casual look at the titles of grants funded most recently by the National Science Foundation’s (NSF’s) Division of Chemical, Bioengineering, Environmental, and Transport Systems, where many academic chemical engineers find at least some support for their research. Key phrases for funded proposals include “dynamic covalent junctions on block copolymer and network self-assembly,” “sustainably derived high-performance nanofiltration membranes,” “ultrahigh-resolution magnetic resonance spectroscopic imaging for label-free molecular imaging,” and “photonic resonator hybrids.” Funding of chemical engineers by the National Institutes of Health is of course even further disconnected from the traditional focus of the undergraduate curriculum.

Federal funding models are also driving a change in the way research is done. Through the 1990s, most research proposals in chemical engineering were written by individual principal investigators and generally were based on elements of intellectual curiosity balanced perhaps with a desire to solve a practical problem. Over the past 20 years, NSF in particular has increased the number of multi-investigator awards relative to single–principal investigator awards. Multiple–principal investigator proposals are often submitted in response to a call for proposals in a particular area (bioseparations or cybersecurity, for example). The data (Figure 2-2) show that this shift has been slow but steady. The impact on academic research has been positive in the sense that there are more interdisciplinary teams, and students are more likely to have multiple advisors or mentors, both of which improve educational outcomes. However, this shift also has led to a decline in funding for basic research by individuals on problems they find interesting. The new model supports innovation, frequently done by large teams, but not necessarily invention, which often comes from individuals (e.g., Wu et al., 2019). In addition to this shift toward large teams, the committee’s sense is that funding opportunities are more prescriptive, or targeted, and less open-ended than in the past. In this context, funding agencies will need to consider the appropriate balance to meet their respective missions.

Suggested Citation:"2 Chemical Engineering Today." National Academies of Sciences, Engineering, and Medicine. 2022. New Directions for Chemical Engineering. Washington, DC: The National Academies Press. doi: 10.17226/26342.
×
Image
FIGURE 2-2 Number of new research projects funded by the National Science Foundation with a single principal investigator (SPI) (light blue; left) or multiple principal investigators (MPI) (dark blue; right) over the past 10 years. The ratio of MPI to SPI awards (solid black line with dashed black trend line) has increased. Data from NSF (2020).

This challenge is not unique to chemical engineering. Tension between basic and applied research characterizes most fields of science and engineering. Inventions cannot be planned and do not arrive on a schedule. Basic research is messy, nonlinear, and expensive. New knowledge relies on serendipity and the preparation of a fertile mind, and not all individuals have the ability to invent. The outcomes of basic research can fundamentally change the course of history, but for that to happen, a period of innovation after invention is usually required. Such innovation is often achieved through applied research or engineering.

On the industrial research front, the linear model of research, development, and commercialization has also been disrupted in the years since the Amundson Report was released. In the past, major companies deployed research and development (R&D) centers in support of their long-term growth objectives. Projects would then be transitioned to the appropriate business areas for commercialization, which often led to major new products and services. The R&D centers also served as a source of technical talent for the business areas because the research innovators would often transition to a business unit along with their technology developments. R&D centers maintained close connections to universities and academic programs as well. The overall system was a path from basic discovery at the university out to the market.

Priorities began to shift in the 1980s. Companies need to create new products or processes to maintain profitability. Large companies in the chemical and oil and gas industries have shifted their research activities toward supporting existing businesses. Companies have set shorter-term goals for corporate research, and their support for university research now focuses more on basic science and has grown only slightly since the 1980s

Suggested Citation:"2 Chemical Engineering Today." National Academies of Sciences, Engineering, and Medicine. 2022. New Directions for Chemical Engineering. Washington, DC: The National Academies Press. doi: 10.17226/26342.
×

(NSB, 2020). These changes have further widened the “valley of death” between discovery and commercialization. To fill that gap, startups, innovation incubators, and many other intermediate models have emerged in and near universities, but the bidirectional connection of research and the marketplace is still challenged. The exception is the pharmaceutical industry, where fundamental drug discovery is still carried out. A comparison of R&D expenditures as a function of total revenue or total chemical sales (Figure 2-3) shows the marked difference between the pharmaceutical industry and the chemical and petroleum industry, respectively.

In a relatively recent development, companies have been expanding or replacing parts of their R&D effort with open innovation models in which they are seeking solutions from outside their R&D divisions or even their companies. Companies are also forming alliances with other companies, universities, and/or national laboratories to solve complex problems, such as the “end of plastic waste” (the Alliance to End Plastic Waste includes more than 40 companies that are pooling funding and expertise to address the problem of plastic waste).

Education and research have historically been viewed as separate activities. Today, there is an opportunity to design new models whereby learning and innovation are connected through new forms of public–private partnerships. Many opportunities exist to build a connected model in which students can take the latest ideas and technology to the marketplace, have experience with a business operation, and return with a perspective on the needs of the market to inspire new research initiatives. Many of the component steps for such a model exist, but new cross-sector collaborations to support and accelerate the growth of this model are needed.

Image
FIGURE 2-3 Research and development (R&D) expenditures in 2020 by the top 10 U.S.-based pharmaceutical (light blue circles) and chemical (dark blue triangles) companies (for which data were available), each compared with their revenue or chemical sales, respectively. Data from Buntz (2021); C&EN (2021); Macrotrends (2021a,b).
Suggested Citation:"2 Chemical Engineering Today." National Academies of Sciences, Engineering, and Medicine. 2022. New Directions for Chemical Engineering. Washington, DC: The National Academies Press. doi: 10.17226/26342.
×

GROWTH OF INTERDISCIPLINARY WORK

The growth of interdisciplinary research reflects the increasing complexity of the problems to be solved and the growing sophistication of the tools needed to solve them. Nonetheless, such research also poses a danger. By definition, one cannot have interdisciplinary work without disciplines. The field of chemical engineering needs to recognize that its graduates are valued for the disciplinary skills they can bring to bear in working with others on a problem. Thus it is important to continue to educate students in the basic skills of chemical engineering, albeit with examples less reflective of an olefin-based business. At the same time, the field needs to be open to the influx of faculty members and practitioners who have not had a traditional chemical engineering education. Chemical engineering has benefited enormously from the influx of mathematicians, physical and other chemists, physicists and materials scientists, biologists, and others to its faculties over the years, a phenomenon expected to continue. One revealing example is that in the United States, chemical engineering is home to most programs in polymer science, a field largely founded in but not embraced by chemistry. This is no time for stasis, but instead a time for the field to expand and grow at its current frontiers while remaining true to the core that defines it. As societal challenges become increasingly complex, science and engineering solutions will necessarily come from connections across disciplines, and the boundaries between disciplines will continue to blur. To contribute to solutions for societal challenges in the coming decades, chemical engineers will need to become increasingly comfortable working across disciplines and as members of interdisciplinary teams. Indeed, this report highlights many areas in which chemical engineers will benefit from interdisciplinary collaborations.

Suggested Citation:"2 Chemical Engineering Today." National Academies of Sciences, Engineering, and Medicine. 2022. New Directions for Chemical Engineering. Washington, DC: The National Academies Press. doi: 10.17226/26342.
×
Page 17
Suggested Citation:"2 Chemical Engineering Today." National Academies of Sciences, Engineering, and Medicine. 2022. New Directions for Chemical Engineering. Washington, DC: The National Academies Press. doi: 10.17226/26342.
×
Page 18
Suggested Citation:"2 Chemical Engineering Today." National Academies of Sciences, Engineering, and Medicine. 2022. New Directions for Chemical Engineering. Washington, DC: The National Academies Press. doi: 10.17226/26342.
×
Page 19
Suggested Citation:"2 Chemical Engineering Today." National Academies of Sciences, Engineering, and Medicine. 2022. New Directions for Chemical Engineering. Washington, DC: The National Academies Press. doi: 10.17226/26342.
×
Page 20
Suggested Citation:"2 Chemical Engineering Today." National Academies of Sciences, Engineering, and Medicine. 2022. New Directions for Chemical Engineering. Washington, DC: The National Academies Press. doi: 10.17226/26342.
×
Page 21
Suggested Citation:"2 Chemical Engineering Today." National Academies of Sciences, Engineering, and Medicine. 2022. New Directions for Chemical Engineering. Washington, DC: The National Academies Press. doi: 10.17226/26342.
×
Page 22
Suggested Citation:"2 Chemical Engineering Today." National Academies of Sciences, Engineering, and Medicine. 2022. New Directions for Chemical Engineering. Washington, DC: The National Academies Press. doi: 10.17226/26342.
×
Page 23
Suggested Citation:"2 Chemical Engineering Today." National Academies of Sciences, Engineering, and Medicine. 2022. New Directions for Chemical Engineering. Washington, DC: The National Academies Press. doi: 10.17226/26342.
×
Page 24
Suggested Citation:"2 Chemical Engineering Today." National Academies of Sciences, Engineering, and Medicine. 2022. New Directions for Chemical Engineering. Washington, DC: The National Academies Press. doi: 10.17226/26342.
×
Page 25
Suggested Citation:"2 Chemical Engineering Today." National Academies of Sciences, Engineering, and Medicine. 2022. New Directions for Chemical Engineering. Washington, DC: The National Academies Press. doi: 10.17226/26342.
×
Page 26
Suggested Citation:"2 Chemical Engineering Today." National Academies of Sciences, Engineering, and Medicine. 2022. New Directions for Chemical Engineering. Washington, DC: The National Academies Press. doi: 10.17226/26342.
×
Page 27
Suggested Citation:"2 Chemical Engineering Today." National Academies of Sciences, Engineering, and Medicine. 2022. New Directions for Chemical Engineering. Washington, DC: The National Academies Press. doi: 10.17226/26342.
×
Page 28
Suggested Citation:"2 Chemical Engineering Today." National Academies of Sciences, Engineering, and Medicine. 2022. New Directions for Chemical Engineering. Washington, DC: The National Academies Press. doi: 10.17226/26342.
×
Page 29
Suggested Citation:"2 Chemical Engineering Today." National Academies of Sciences, Engineering, and Medicine. 2022. New Directions for Chemical Engineering. Washington, DC: The National Academies Press. doi: 10.17226/26342.
×
Page 30
Suggested Citation:"2 Chemical Engineering Today." National Academies of Sciences, Engineering, and Medicine. 2022. New Directions for Chemical Engineering. Washington, DC: The National Academies Press. doi: 10.17226/26342.
×
Page 31
Next: 3 Decarbonization of Energy Systems »
New Directions for Chemical Engineering Get This Book
×
 New Directions for Chemical Engineering
Buy Paperback | $50.00 Buy Ebook | $40.99
MyNAP members save 10% online.
Login or Register to save!
Download Free PDF

Over the past century, the work of chemical engineers has helped transform societies and the lives of individuals, from the synthetic fertilizers that helped feed the world to the development of novel materials used in fuels, electronics, medical devices, and other products. Chemical engineers' ability to apply systems-level thinking from molecular to manufacturing scales uniquely positions them to address today’s most pressing problems, including climate change and the overuse of resources by a growing population.

New Directions for Chemical Engineering details a vision to guide chemical engineering research, innovation, and education over the next few decades. This report calls for new investments in U.S. chemical engineering and the interdisciplinary, cross-sector collaborations necessary to advance the societal goals of transitioning to a low-carbon energy system, ensuring our production and use of food and water is sustainable, developing medical advances and engineering solutions to health equity, and manufacturing with less waste and pollution. The report also calls for changes in chemical engineering education to ensure the next generation of chemical engineers is more diverse and equipped with the skills necessary to address the challenges ahead.

READ FREE ONLINE

  1. ×

    Welcome to OpenBook!

    You're looking at OpenBook, NAP.edu's online reading room since 1999. Based on feedback from you, our users, we've made some improvements that make it easier than ever to read thousands of publications on our website.

    Do you want to take a quick tour of the OpenBook's features?

    No Thanks Take a Tour »
  2. ×

    Show this book's table of contents, where you can jump to any chapter by name.

    « Back Next »
  3. ×

    ...or use these buttons to go back to the previous chapter or skip to the next one.

    « Back Next »
  4. ×

    Jump up to the previous page or down to the next one. Also, you can type in a page number and press Enter to go directly to that page in the book.

    « Back Next »
  5. ×

    Switch between the Original Pages, where you can read the report as it appeared in print, and Text Pages for the web version, where you can highlight and search the text.

    « Back Next »
  6. ×

    To search the entire text of this book, type in your search term here and press Enter.

    « Back Next »
  7. ×

    Share a link to this book page on your preferred social network or via email.

    « Back Next »
  8. ×

    View our suggested citation for this chapter.

    « Back Next »
  9. ×

    Ready to take your reading offline? Click here to buy this book in print or download it as a free PDF, if available.

    « Back Next »
Stay Connected!