Pharmaceutical Century Patents & Potions (1800's to 1919) We live today in a world of drugs. Drugs for pain, drugs for disease, drugs for allergies, drugs for pleasure, and drugs for mental health. Drugs that have been rationally designed; drugs that have been synthesized in the factory or purified from nature. Drugs fermented and drugs engineered. Drugs that have been clinically tested. Drugs that, for the most part, actually do what they are supposed to. Effectively. Safely. By no means was it always so. Before the end of the 19th century, medicines were concocted with a mixture of empiricism and prayer. Trial and error, inherited lore, or mystical theories were the basis of the world’s pharmacopoeias. The technology of making drugs was crude at best: Tinctures, poultices, soups, and teas were made with water- or alcohol-based extracts of freshly ground or dried herbs or animal products such as bone, fat, or even pearls, and sometimes from minerals best left in the ground—mercury among the favored. The difference between a poison and a medicine was a hazy differentiation at best: In the 16th century, Paracelsus declared that the only difference between a medicine and a poison was in the dose. All medicines were toxic. It was cure or kill. “Rational treatments” and “rational drug design” of the era were based on either the doctrine of humors (a pseudoastrological form of alchemical medicine oriented to the fluids of the body: blood, phlegm and black and yellow bile) or the doctrine of signatures. (If a plant looks like a particular body part, it must be designed by nature to influence that part. Lungwort, for example, was considered good for lung complaints by theorists of the time because of its lobe-shaped leaves.) Neither theory, as might be expected, guaranteed much chance of a cure. Doctors and medicines were popular, despite their failures. As pointed out by noted medical historian Charles E.Rosenberg, a good bedside manner and a dose of something soothing (or even nasty) reassured the patient that something was being done, that the disease was not being ignored. Blood and mercury By the first part of the 19th century, the roots of modern pharmacy had taken hold with a wave of heroic medicine. Diseases were identified by symptom, and attacking the symptom as vigorously as possible was the high road to health. Bloodletting dominated the surgeon’s art, and dosing patients with powerful purgatives and cathartics became the order of the day in an attempt to match the power of the disease with the power of the drug. Bleed them till they faint. (It is difficult to sustain a raging fever or pounding pulse when there is too little blood in the body, so the symptoms, if not what we would call the disease, seemed to vanish.) Dose them with calomel till they drool and vomit. (Animals were thought to naturally expel toxins this way.) Cleanse both stomach and bowels violently to remove the poisons there. Certainly these methods were neither pleasant nor very effective at curing patients already weakened by disease. George Washington died in misery from bloodletting; Abraham Lincoln suffered chronic mercury poisoning and crippling constipation from his constant doses of “blue mass.” The “cure” was, all too often, worse than the disease. In the second half of the 19th century, things changed remarkably as the industrial revolution brought technological development to manufacturing and agriculture and inspired the development of medical technology. Spurred in part by a reaction against doctors and their toxic nostrums, patent medicines and in particular homeopathy (which used extreme dilutions of otherwise toxic compounds) became popular and provided an “antidote” to the heroic treatments of the past. Not helpful, but at least harmless for the most part, these new drugs became the foundation of a commodity-based medicine industry that galvanized pharmacist and consumer alike. Technology entered in the form of pill and powder and potion making. Almost by accident, a few authentic drugs based on the wisdom and herbal lore of the past were developed: quinine, digitalis, and cocaine. Ultimately, these successes launched the truly modern era. The century ended with the development of the first of two synthesized drugs that represent the triumph of chemistry over folklore and technology over cookery. The development of antipyrine in 1883 and aspirin in 1897 set the stage for the next 10 decades of what we can look back on in retrospect as the Pharmaceutical Century. With new knowledge of microbial pathogens and the burgeoning wisdom of vaccine technology, the first tentative steps were taken to transform medicines to a truly scientific foundation. From these scattered seeds, drug technology experienced remarkable if chaotic growth in the first two decades of the 20th century, a period that can be likened to a weedy flowering of quackery and patent medicines twining about a hardening strand of authentic science and institutions to protect and nourish it. Staging the Pharmaceutical Century In the latter half of the 19th century, numerous beneficent botanicals took center stage in the world’s pharmacopoeias. Cocaine was first extracted from coca leaves in 1860; salicylic acid—the forerunner of aspirin—was extracted from willow bark in 1874 for use as a painkiller. Quinine and other alkaloids had long been extracted from China bark; but an antifebrile subcomponent, quinoline, was not synthesized in the lab until 1883 by Ludwig Knorr. The first truly synthetic pain reliever, antipyrine, was produced from quinoline derivatives. Digitalis from foxglove and strophantin from an African dogbane were both botanicals purified for use against heart disease. The opium poppy provided a wealth of pain relievers: opium, morphine, codeine, and heroin. But it was not until the birth of medical microbiology that the true breakthroughs occurred, and science—rather than empiricism—took center stage in the development of pharmaceuticals. Murderous microbes The hallmark of 19th-century medicine has to be the microbial theory of disease. The idea that infectious diseases were caused by microscopic living agents provided an understanding of the causes and the potential cures for ills from anthrax to whooping cough. Technology made the new framework possible. The brilliance of European lens makers and microscopists, coupled with the tinkering of laboratory scientists who developed the technologies of sterilization and the media and methods for growing and staining microbes, provided the foundation of the new medical science that would explode in the 20th century. These technologies offered proof and intelligence concerning the foe against which pharmaceuticals, seen thereafter as weapons of war, could be tested and ultimately designed. In 1861, the same year that the American Civil War began, Ignaz Semmel weis published his research on the transmissible nature of purperal (childbed) fever. His theories of antisepsis were at first vilified by doctors who could not believe their unwashed hands could transfer disease from corpses or dying patients to healthy women. But eventually, with the work of Robert Koch, Joseph Lister, and Louis Pasteur adding proof of the existence and disease-causing abilities of microorganisms, a worldwide search for the microbial villains of a host of historically deadly diseases began. In 1879, as part of the new “technology,” Bacterium coli was discovered (it was renamed Escherichia after its discoverer, Theodor Escherich, in 1919). It quickly became the quintessential example of an easily grown, “safe” bacteria for laboratory practice. New growth media, new sterile techniques, and new means of isolating and staining bacteria rapidly developed. The ability to grow “pathogens” in culture proved remarkably useful. Working with pure cultures of the diphtheria bacillus in Pasteur’s laboratory in 1888, Emile Roux and Alexandre Yersin first isolated the deadly toxin that causes most of diphtheria’s lethal effects. One by one over the next several decades, various diseases revealed their microbial culprits to the so-called microbe-hunters. Initially, most American physicians were loath to buy into germ theory, seeing it as a European phenomenon incompatible with the “truth” of spontaneous generation and as a threat to the general practitioner from the growing cadre of scientifically trained laboratory microbiologists and specialist physicians. “Anti-contagionists” such as the flamboyant Colonel George E. Waring Jr., pamphleteer, consulting engineer, and phenomenally effective warrior in the sanitation movement, ultimately held sway. Filth was considered the source of disease. A host of sewage projects, street-cleaning regimens, and clean water systems swept urban areas across the United States, with obvious benefits. Ultimately, the germ theory of infectious diseases had to be accepted, especially as the theoretical foundation behind the success of the sanitation movement. And with the production of vaccines and antitoxins, older medical frameworks fell by the wayside, though rural American physicians were still recommending bleeding and purgatives as cures well into the first few decades of the 20th century. Victorious vaccines The most significant outgrowth of the new germ theory, and the one that created the greatest demand for new technologies for implementation, was the identification and production of the new “immunologicals”—drugs that are, in essence, partially purified components or fractions of animal blood. In 1885, Pasteur developed attenuated rabies vaccine—a safe source of “active” immunity (immunity developed against a form or component of the disease-causing microorganism by the body’s own immune system). Vaccines would be developed against a variety of microorganisms in rapid succession over the next several decades. But active immunity was perhaps not the most impressive result of the immunologicals. Antitoxins (antibodies isolated against disease organisms and their toxins from treated animals), when injected into infected individuals, provided salvation from otherwise fatal diseases. This technology began in 1890 when Emil von Behring and Shibasaburo Kitasato isolated the first antibodies against tetanus and, soon after, diphtheria. In 1892, Hoechst Pharma developed a tuberculin antitoxin. These vaccines and antitoxins would form the basis of a new pharmaceutical industry. Perhaps as important as the development of these new immunologics was the impetus toward standardization and testing that a new generation of scientist-practitioners such as Koch and Pasteur inspired. These scientists’ credibility and success rested upon stringent control—and ultimately, government regulation—of the new medicines. Several major institutions sprang up in Europe and the United States to manufacture and/or inspect in bulk the high volume of vaccines and antitoxins demanded by a desperate public suddenly promised new hope against lethal diseases. These early controls helped provide a bulwark against contamination and abuse. Such control would not be available to the new synthetics soon to dominate the scene with the dawn of “scientific” chemotherapy. Medicinal chemistry Parallel (and eventually linked) to developments in biology, the chemist’s art precipitously entered the medicinal arena in 1856 when Englishman William Perkin, in an abortive attempt to synthesize quinine, stumbled upon mauve, the first synthesized coal tar dye. This discovery led to the development of many synthetic dyes but also to the realization that some of these dyes had therapeutic effects. Synthetic dyes, and especially their medicinal “side effects,” helped put Germany and Switzerland in the forefront of both organic chemistry and synthesized drugs. The dye–drug connection was a two-way street: The antifever drug Antifebrin, for example, was derived from aniline dye in 1886. The chemical technology of organic synthesis and analysis seemed to offer for the first time the potential to scientifically ground the healer’s art in a way far different from the “cookery” of ancient practitioners. In 1887, phenacetin, a pain reliever, was developed by Bayer specifically from synthetic drug discovery research. The drug eventually fell into disfavor because of its side effect of kidney damage. Ten years later, also at Bayer, Felix Hoffman synthesized acetylsalicylic acid (aspirin). First marketed in 1899, aspirin has remained the most widely used of all the synthetics. Many other new technologies also enhanced the possibilities for drug development and delivery. The advent of the clinical thermometer in 1870 spearheaded standardized testing and the development of the antifever drugs. In 1872, Wyeth invented the rotary tablet press, which was critical to the mass marketing of drugs. By 1883, a factory was producing the first commercial drug (antipyrine) in a ready-dosaged, prepackaged form. With the discovery of X-rays in 1895, the first step was taken toward X-ray crystallography, which would become the ultimate arbiter of complex molecular structure, including proteins and dna. The Pharmaceutical Century Not only did the early 1900s bring the triumph of aspirin as an inexpensive and universal pain reliever—the first of its kind—but the science of medicine exploded with a new understanding of the human body and its systems. Although not immediately translated into drugs, these discoveries would rapidly lead to a host of new pharmaceuticals and a new appreciation of nutrition as a biochemical process and hence a potential source of drugs and drug intervention. Of equal if not more importance to the adoption and implementation of the new technologies was the rise of public indignation—a demand for safety in food and medicines that began in Europe and rapidly spread to the United States. Tainted food and public furor The developing understanding of germ theory and the increasing availability of immunologics and chemical nostrums forced recognition that sanitation and standardization were necessary for public health and safety. First in Europe, and then in the United States, the new technologies led to the growth of new public and semipublic institutions dedicated to producing and/or assaying the effectiveness and safety of pharmaceuticals and foods in addition to those dedicated to sanitation and public disease control. Unfortunately, the prevalence of disease among the poor created a new line of prejudice against these presumed “unsanitary” subclasses. In the United States, where the popular sanitation movement could now be grounded in germ theory, this fear of contagion manifested among the developing middle classes was directed especially against immigrants—called “human garbage” by pundits such as American social critic Herbert George in 1883. This led to the Immigration Act of 1891, which mandated physical inspection of immigrants for diseases of mind and body—any number of which could be considered cause for quarantine or exclusion. Also in 1891, the Hygienic Laboratory (founded in 1887 and the forerunner of the National Institutes of Health) moved from Staten Island (New York City) to Washington, DC—a sign of its growing importance. That same year, the first International Sanitary Convention was established. Although restricted to efforts to control and prevent cholera, it would provide a model of things to come in the public health arena. In 1902, an International Sanitary Bureau (later renamed the Pan American Sanitary Bureau and then the Pan American Sanitary Organization) was established in Washington, DC, and became the forerunner of today’s Pan American Health Organization, which also serves as the World Health Organization’s Regional Office for the Americas. Fears of contagion on the one hand and poisoning on the other, resulting from improperly prepared or stored medicines, led to the 1902 Biologicals Controls Act, which regulates the interstate sale of viruses, serums, antitoxins, and similar products. One of the significant outgrowths of the new “progressive” approach to solving public health problems with technological expertise and government intervention was the popularity and influence of a new class of journalists known as the Muckrakers. Under their impetus, and as the result of numerous health scandals, The 1906 U.S. Pure Food and Drugs Act, after years of planning by U.S. Department of Agriculture (USDA) researchers such as John Wiley, was passed easily. The act established the USDA’s Bureau of Chemistry as the regulatory agency. Unfortunately, the act gave the federal government only limited powers of inspection and control over the industry. Many patent medicines survived this first round of regulation. The American Medical Association (AMA) created a Council on Pharmacy and Chemistry to examine the issue and then established a chemistry laboratory to lead the attack on the trade in patent medicines that the Pure Food and Drugs Act had failed to curb. The ama also published New and Nonofficial Remedies annually in an effort to control drugs by highlighting serious issues of safety and inefficacy. This publication prompted rapid changes in industry standards. International health procedures continued to be formalized also—L’Office International d’Hygiène Publique (OIHP) was established in Paris in 1907, with a permanent secretariat and a permanent committee of senior public health officials. Military and geopolitical concerns would also dominate world health issues. In 1906, the Yellow Fever Commission was established in Panama to help with U.S. efforts to build the canal; in 1909, the U.S. Army began mass vaccination against typhoid. Nongovernmental organizations also rallied to the cause of medical progress and reform. In 1904, for example, the U.S. National Tuberculosis Society was founded (based on earlier European models) to promote research and social change. It was one of many groups that throughout the 20th century were responsible for much of the demand for new medical technologies to treat individual diseases. Grassroots movements such as these flourished. Public support was often behind the causes. In 1907, Red Cross volunteer Emily Bissell designed the first U.S. Christmas Seals (the idea started in Denmark). The successful campaign provided income to the Tuberculosis Society and a reminder to the general public of the importance of medical care. Increased public awareness of diseases and new technologies such as vaccination, antitoxins, and later, “magic bullets,” enhanced a general public hunger for new cures. The movement into medicine of government and semipublic organizations such as the AMA and the Tuberculosis Society throughout the latter half of the 19th and beginning of the 20th centuries set the stage for a new kind of medicine that was regulated, tested, and “public.” Combined with developments in technology and analysis that made regulation possible, public scrutiny slowly forced medicine to come out from behind the veil of secret nostrums and alchemical mysteries. The crowning of chemistry It was not easy for organized science, especially chemistry, to take hold in the pharmaceutical realm. Breakthroughs in organic synthesis and analysis had to be matched with developments in biochemistry, enzymology, and general biology. Finally, new medicines could be tested for efficacy in a controlled fashion using new technologies—laboratory animals, bacterial cultures, chemical analysis, clinical thermometers, and clinical trials, to name a few. Old medicines could be debunked using the same methods—with public and nongovernmental organizations such as the ama providing impetus. At long last, the scientific community began to break through the fog of invalid information and medical chicanery to attempt to create a new pharmacy of pharmaceuticals based on chemistry, not caprice. The flowering of biochemistry in the early part of the new century was key, especially as it related to human nutrition, anatomy, and disease. Some critical breakthroughs in metabolic medicine had been made in the 1890s, but they were exceptions rather than regular occurrences. In 1891, myedema was treated with sheep thyroid injections. This was the first proof that animal gland solutions could benefit humans. In 1896, Addison’s disease was treated with chopped up adrenal glands from a pig. These test treatments provided the starting point for all hormone research. Also in 1891, a pair of agricultural scientists developed the Atwater–Rosa calorimeter for large animals. Ultimately, it provided critical baselines for human and animal nutrition studies. But it wasn’t until the turn of the century that metabolic and nutritional studies truly took off. In 1900, Karl Landsteiner discovered the first human blood groups: O, A, and B. That same year, Frederick Hopkins discovered tryptophan and demonstrated in rat experiments that it was an “essential” amino acid—the first discovered. In 1901, fats were artificially hydrogenated for storage for the first time (providing a future century of heart disease risk). Eugene L. Opie discovered the relationship of islets of Langerhans to diabetes mellitus, thus providing the necessary prelude to the discovery of insulin. Japanese chemist Jokichi Takamine isolated pure epinephrine (adrenaline). And E. Wildiers discovered “a new substance indispensable for the development of yeast.” Growth substances such as this eventually became known as vitamines and later, vitamins. In 1902, proteins were first shown to be polypeptides, and the AB blood group was discovered. In 1904, the first organic coenzyme—cozymase—was discovered. In 1905, allergies were first described as a reaction to foreign proteins by Clemens von Pirquet, and the word “hormone” was coined. In 1906, Mikhail Tswett developed the all-important technique of column chromatography. In 1907, Ross Harrison developed the first animal cell culture using frog embryo tissues. In 1908, the first biological audioradiograph was made—of a frog. In 1909, Harvey Cushing demonstrated the link of pituitary hormone to giantism. Almost immediately after Svante August Arrhenius and Soren Sorensen demonstrated in 1909 that pH could be measured, Sorenson pointed out that pH can affect enzymes. This discovery was a critical step in the development of a biochemical model of metabolism and kinetics. So many breakthroughs of medical significance occurred in organic chemistry and biochemistry in the first decade of the Pharmaceutical Century that no list can do more than scratch the surface. Making magic bullets It was not the nascent field of genetics, but rather a maturing chemistry that would launch the most significant early triumph of the Pharmaceutical Century. Paul Ehrlich first came up with the magic bullet concept in 1906. (Significant to the first magic bullet’s ultimate use, in this same year, August von Wasserman developed his syphilis test only a year after the bacterial cause was determined.) However, it wasn’t until 1910 that Ehrlich’s arsenic compound 606, marketed by Hoechst as Salvarsan, became the first effective treatment for syphilis. It was the birth of chemotherapy. With the cure identified and the public increasingly aware of the subject, it was not surprising that the “progressive” U.S. government intervened in the public health issue of venereal disease. The Charmerlain–Kahn Act of 1918 provided the first federal funding specifically designated for controlling venereal disease. It should also not be a surprise that this attack on venereal disease came in the midst of a major war. Similar campaigns would be remounted in the 1940s. The “fall” of chemotherapy Salvarsan provided both the promise and the peril of chemotherapy. The arsenicals, unlike the immunologicals, were not rigidly controlled and were far more subject to misprescription and misuse. (They had to be administered in an era when injection meant opening a vein and percolating the solution into the bloodstream through glass or rubber tubes.) The problems were almost insurmountable, especially for rural practitioners. The toxicity of these therapeutics and the dangers associated with using them became their downfall. Most clinicians of the time thought the future was in immunotherapy rather than chemotherapy, and it wasn’t until the antibiotic revolution of the 1940s that the balance would shift. Ultimately, despite the manifold breakthroughs in biochemistry and medicine, the end of the ’Teens was not a particularly good time for medicine. The influenza pandemic of 1918–1920 clearly demonstrated the inability of medical science to stand up against disease. More than 20 million people worldwide were killed by a flu that attacked not the old and frail but the young and strong. This was a disease that no magic bullet could cure and no government could stamp out. Both war and pestilence set the stage for the Roaring Twenties, when many people were inclined to “eat, drink, and make merry” as if to celebrate the optimism of a world ostensibly at peace. Still, a burgeoning science of medicine promised a world of wonders yet to come. Technological optimism and industrial expansion provided an antidote to the malaise caused by failed promises revealed in the first two decades of the new century. But even these promises were suspect as the Progressive Era drew to a close. Monopoly capitalism and renewed conservatism battled against government intervention in health care as much as the economy did and became a familiar refrain. The continued explosive growth of cities obviated many of the earlier benefits in sanitation and hygiene with a host of new “imported” diseases. The constitutive bad health and nutrition of both the urban and rural poor around the world grew worse with the economic fallout of the war. Many people were convinced that things would only get worse before they got better. The Pharmaceutical Century had barely begun. SALVING WITH SCIENCE (1920s & 1930s) Throughout the 1920s and 1930s, new technologies and new science intersected as physiology led to the discovery of vitamins and to increasing knowledge of hormones and body chemistry. New drugs and new vaccines flowed from developments started in the previous decades. Sulfa drugs became the first of the anti bacterial wonder drugs promising broad-spectrum cures. Penicillin was discovered, but its development had to await new technology (and World War II, which hastened it). New instruments such as the ultracentrifuge and refined techniques of X-ray crystallography paralleled the development of virology as a science. Isoelectric precipitation and electrophoresis first became important for drug purification and analysis. Overall, the medical sciences were on a firmer footing than ever before. This was the period in which the Food and Drug Administration (FDA) gained independence as a regulatory agency.And researchers reveled in the expanding knowledge base and their new instruments. They created a fusion of medicine and machines that would ultimately be known as “molecular biology.” The assault on disease Just as pharmaceutical chemists sought “magic bullets” for myriad diseases in the first two decades of the century, chemists inthe 1920s and 1930s expanded the search for solutions to the bacterial and viral infections that continued to plague humankind. Yet, the first great pharmaceutical discovery of the 1920s addressed not an infectious disease but a physiological disorder. Diabetes mellitus is caused by a malfunction of the pancreas, resulting in the failure of that gland to produce insulin, the hormone that regulates blood sugar. For most of human history, this condition meant certain death. Since the late 19th century, when the connection between diabetes and the pancreas was first determined, scientists had attempted to isolate the essential hormone and inject it into the body to control the disorder. Using dogs as experimental subjects, numerous researchers had tried and failed, but in 1921, Canadian physician Frederick Banting made the necessary breakthrough. Banting surmised that if he tied off the duct to the pancreas of a living dog and waited until the gland atrophied before removing it, there would be no digestive juices left to dissolve the hormone, which was first called iletin. Beginning in the late spring of 1921, Banting worked on his project at the University of Toronto with his assistant, medical student Charles Best. After many failures, one of the dogs whose pancreas had been tied off showed signs of diabetes. Banting and Best removed the pancreas, ground it up, and dissolved it in a salt solution to create the long-sought extract. They injected the extract into the diabetic dog, and within a few hours the canine’s blood sugar returned to normal. The scientists had created the first effective treatment for diabetes. At the insistence of John Macleod, a physiologist at the University of Toronto who provided the facilities for Banting’s work, biochemists James Collip and E. C. Noble joined the research team to help purify and standardize the hormone, which was renamed insulin. Collip purified the extract for use in human subjects, and enough successful tests were performed on diabetic patients to determine that the disorder could be reversed. The Connaught Laboratories in Canada and the Eli Lilly Co. in the United States were awarded the rights to manufacture the drug. Within a few years, enough insulin was being produced to meet the needs of diabetics around the world. Although Banting and Best had discovered the solution to a problem that had troubled humans for millennia, it was Lilly’s technical developments (such as the use of an isoelectric precipitation step) that enabled large-scale collection of raw material, extraction and purification of insulin, and supplying of the drug in a state suitable for clinical use. Only after proper bulk production and/or synthesis techniques were established did insulin and many other hormones discovered in the 1920s and 1930s (such as estrogen, the corticosteroids, and testosterone) become useful and readily available to the public. This would continue to be the case with most pharmaceutical breakthroughs throughout the century. Although the isolation and development of insulin was a critically important pharmaceutical event, diabetes was by no means the greatest killer of the 19th and early 20th centuries. That sordid honor belonged to infectious diseases, especially pneumonia, and scientists in the 1920s and 1930s turned with increasing success to the treatment of some of the most tenacious pestilences. Paul Ehrlich introduced the world to chemotherapy in the early years of the 20th century, and his successful assault on syphilis inspired other chemists to seek “miracle drugs” and “magic bullets.” But what was needed was a drug that could cure general bacterial infections such as pneumonia and septicemia. Bacteriologists began in the 1920s to experiment with dyes that were used to stain bacteria to make them more visible under the microscope, and a breakthrough was achieved in the mid-1930s. Sulfa drugs and more The anti-infective breakthrough occurred at Germany’s I. G. Farben, which had hired Gerhard Domagk in the late 1920s to direct its experimental pathology laboratory in a drive to become a world leader in the production of new drugs. Domagk performed a series of experiments on mice infected with streptococcus bacteria. He discovered that some previously successful compounds killed the bacteria in mice but were too toxic to give to humans. In 1935, after years of experimentation, Domagk injected an orange-red azo dye called Prontosil into a group of infected mice. The dye, which was primarily used to color animal fibers, killed the bacteria, and, most importantly, all the mice survived. The first successful use of Prontosil on humans occurred weeks later, when Domagk gave the drug to a desperate doctor treating an infant dying of bacterial infection. The baby lived, but this did not completely convince the scientific community of the drug’s efficacy. Only when 26 women similarly afflicted with life-threatening infections were cured during clinical trials in London in late 1935 did Prontosil become widely known and celebrated for its curative powers. The active part of Prontosil was a substance called sulfanilamide, so termed by Daniele Bovet of the Pasteur Institute, who determined that Prontosil broke down in the body and that only a fragment of the drug’s molecule worked against an infection. After the discovery of the active ingredient, more than 5000 different “sulfa” drugs were made and tested, although only about 15 ultimately proved to be of value. In 1939, Domagk received the 1939 Nobel Prize in Physiology or Medicine. Sulfanilamide was brought to the United States by Perrin H. Long and Eleanor A. Bliss, who used it in clinical applications at Johns Hopkins University in 1936. It was later discovered that the sulfa drugs, or sulfonamides, did not actually kill bacteria outright, like older antiseptics, but halted the growth and multiplication of the bacteria, while the body’s natural defenses did most of the work. Certainly the most famous antibacterial discovered in the 1920s and 1930s was penicillin—which was found through almost sheer serendipity. In the years after World War I, Alexander Fleming was seeking better antiseptics, and in 1921 he found a substance in mucus that killed bacteria. After further experimentation, he learned that the substance was a protein, which he called lysozyme. Although Fleming never found a way to purify lysozymes or use them to treat infectious diseases, the discovery had implications for his later encounter with penicillin because it demonstrated the existence of substances that are lethal to certain microbes and harmless to human tissue. Fleming’s major discovery came almost seven years later. While cleaning his laboratory one afternoon, he noticed large yellow colonies of mold overgrowing a culture of staphylococcus bacteria on an agar plate. Fleming realized that something was killing the bacteria, and he proceeded to experiment with juice extracted from the mold by spreading it on agar plates covered with more bacteria. He found that even when the juice was highly diluted, it destroyed the bacteria. Calling the new antiseptic penicillin, after the Latin term for brush, Fleming had two assistants purify the mold juice, but he performed no tests on infected animal subjects. He published a paper in 1929 discussing the potential use of penicillin in surgical dressings but went no further. It wasn’t until the 1940s that penicillin was taken up by the medical community. Another important achievement in antibacterial research occurred in the late 1930s, when Rene Dubos and colleagues at the Rockefeller Institute for Medical Research inaugurated a search for soil microorganisms whose enzymes could destroy lethal bacteria. The hope was that the enzymes could be adapted for use in humans. In 1939, Dubos discovered a substance extracted from a soil bacillus that cured mice infected with pneumococci. He named it tyrothricin, and it is regarded as the first antibiotic to be established as a therapeutic substance. The 1920s and 1930s were also interesting times in malaria research. The popular antimalarial drug chloroquine was not formally recognized in the United States until 1946, but it had been synthesized 12 years before at Germany’s Bayer Laboratories under the name Resochin. While much of pharmaceutical science concentrated on finding answers to the problems posed by bacterial infections, there was some significant work done on viruses as well. Viruses were identified in the late 19th century by Dutch botanist Martinus Beijerinck, but a virus was not crystallized until 1935, when biochemist Wendell Stanley processed a ton of infected tobacco leaves down to one tablespoon of crystalline powder—tobacco mosaic virus. Unlike bacteria, viruses proved to be highly resistant to assault by chemotherapy, and thus antiviral research during the era did not yield the successes of antibiotic research. Most clinical research was dedicated to the search for vaccines and prophylactics rather than treatments. Polio was one of the most feared scourges threatening the world’s children in the 1920s and 1930s, but no real breakthroughs came until the late 1940s. Scientifically, the greatest progress in antiviral research was probably made in investigations of the yellow fever virus, which were underwritten by the Rockefeller Foundation in the 1930s. But as in the case of polio, years passed before a vaccine was developed. In the meantime, efforts by the public health services of many nations, including the United States, promoted vaccination as part of the battle against several deadly diseases, from smallpox to typhoid fever. Special efforts were often made to attack the diseases in rural areas, where few doctors were available. Vitamins and deficiency diseases Whereas the medical fight against pernicious bacteria and viruses brings to mind a military battle against invading forces, some diseases are caused by internal treachery and metabolic deficiency rather than external assault. This notion is now commonplace, yet in the early 20th century it was hotly contested. For the first time, scientists of the era isolated and purified “food factors,” or “vitamines,” and understood that the absence of vitamins has various detrimental effects on the body, depending on which ones are in short supply. Vitamins occur in the body in very small concentrations; thus, in these early years, determining which foods contained which vitamins, and analyzing their structure and effects on health, was complex and time-intensive. Ultimately, scientists discerned that vitamins are essential for converting food into energy and are critical to human growth. As a result of this research, by the late 1930s, several vitamins and vitamin mixtures were used for therapeutic purposes. The isolation of specific vitamins began in earnest in the second decade of the 20th century and continued into the 1920s and 1930s. Experiments in 1916 showed that fat-soluble vitamin A was necessary for normal growth in young rats; in 1919, Harry Steenbock, then an agricultural chemist at the University of Michigan, observed that the vitamin A content of vegetables varies with the degree of vegetable pigmentation. It was later determined that vitamin A is derived from the plant pigment carotene. Also in 1919, Edward Mellanby proved that rickets is caused by a dietary deficiency. His research indicated that the deficiency could be overcome—and rickets prevented or cured—by adding certain fats to the diet, particularly cod-liver oil. At first, Mellanby thought that vitamin A was the critical factor, but further experimentation did not support this hypothesis. Three years later, Elmer McCollum and associates at Johns Hopkins University offered clear proof that vitamin A did not prevent rickets and that the antirachitic factor in cod-liver oil was the fat-soluble vitamin D. The research team soon developed a method for estimating the vitamin D content in foods. Experiments on vitamin D continued into the mid-1920s. The most significant were the projects of Steenbock and Alfred Hess, working in Wisconsin and New York, respectively, who reported that antirachitic potency could be conveyed to some biological materials be exposing them to a mercury-vapor lamp. The substance in food that was activated by ultraviolet radiation was not fat, but a compound associated with fat called ergosterol, which is also present in human skin. The scientists surmised that the explanation of the antirachitic effect of sunlight is that ultraviolet rays form vitamin D from ergosterol in the skin, which then passes into the blood. The term vitamin D-1 was applied to the first antirachitic substance to be isolated from irradiated ergosterol, and we now know that there are several forms of vitamin D. Other important vitamin studies took place in the 1920s and 1930s that had implications for the future of pharmaceutical chemistry. In 1929, Henrik Dam, a Danish biochemist, discovered that chicks fed a diet that contained no fat developed a tendency toward hemophilia. Five years later, Dam and colleagues discovered that if hemp seeds were added to the chicks’ diet, bleeding did not occur. The substance in the seeds that protected against hemorrhage was named vitamin K, for koagulation vitamin. In 1935, Armand Quick and colleagues at Marquette University reported that the bleeding often associated with jaundiced patients was caused by a decrease in the blood coagulation factor prothrombin. This study was complemented by a report by H. R. Butt and E. D. Warner that stated that a combination of bile salts and vitamin K effectively relieved the hemorrhagic tendency in jaundiced patients. All of these scientists’ work pointed to the conclusion that vitamin K was linked to the clotting of blood and was necessary for the prevention of hemorrhage, and that vitamin K was essential for the formation of prothrombin. Dam and the Swiss chemist Paul Kirrer reported in 1939 that they had prepared pure vitamin K from green leaves. In the same year, Edward Doisy, a biochemist at Saint Louis University, isolated vitamin K from alfalfa, determined its chemical composition, and synthesized it in the laboratory. Vitamin K was now available for treating patients who suffered from blood clotting problems. Dam and Doisy received the 1943 Nobel Prize in Physiology or Medicine for their work. With the advent of successful research on vitamins came greater commercial exploitation of these substances. In 1933, Tadeus Reichstein synthesized ascorbic acid (vitamin C), making it readily available thereafter. The consumption of vitamins increased in the 1930s, and popular belief held them to be almost magical. Manufacturers, of course, did not hesitate to take advantage of this credulity. There was no informed public regulation of the sale and use of vitamins, and, as some vitamins were dangerous in excess quantities, this had drastic results in isolated cases. Water-soluble vitamins such as vitamin C easily flow out of the body through the kidneys, but the fat-soluble vitamins, such as A, D, and K, could not so easily be disposed of and might therefore prove especially dangerous. Many physicians of the era did nothing to discourage popular misconceptions about vitamins, or harbored relatively uncritical beliefs themselves. The FDA and federal regulation The threat posed by unregulated vitamins was not nearly as dangerous as the potential consequences of unregulated drugs. Yet legislation was enacted only when drug tragedies incensed the public and forced Congress to act. After the 1906 Pure Food and Drug Act, there was no federal legislation dealing with drugs for decades, although the American Medical Association (AMA) did attempt to educate physicians and the public about pharmaceuticals. The AMA published books exposing quack medicines, gradually adopted standards for advertisements in medical journals, and in 1929, initiated a program of testing drugs and granting a Seal of Acceptance to those meeting its standards. Only drugs that received the seal were eligible to advertise in AMA journals. Dangerous drugs were still sold legally, however, because safety testing was not required before marketing. Well into the 1930s, pharmaceutical companies still manufactured many 19th and early 20th century drugs that were sold in bulk to pharmacists, who then compounded them into physicians’ prescriptions. But newer drugs, such as many biologicals and sulfa drugs (after 1935), were packaged for sale directly to consumers and seemed to represent the future of drug manufacturing. In 1937, an American pharmaceutical company produced a liquid sulfa drug. Attempting to make sulfanilamide useful for injections, the company mixed it with diethylene glycol—the toxic chemical now used in automobile antifreeze. Ultimately sold as a syrup called Elixir of Sulfanilamide, the drug concoction was on the market for two months, in which time it killed more than 100 people, including many children, who drank it. Under existing federal legislation, the manufacturer could be held liable only for mislabeling the product. In response to this tragedy and a series of other scandals, Congress passed the Food, Drug and Cosmetic Act of 1938, which banned drugs that were dangerous when used as directed, and required drug labels to include directions for use and appropriate warnings. The act also required new drugs to be tested for safety before being granted federal government approval and created a new category of drugs that could be dispensed to a patient only at the request of a physician. Before the act was passed, patients could purchase any drug, except narcotics, from pharmacists. The Food and Drug Administration (the regulatory division established in 1927 from the former Bureau of Chemistry) was given responsibility for implementing these laws. The 1938 legislation is the basic law that still regulates the pharmaceutical industry. New manufacturing and mass-marketing methods demanded changes in federal oversight, because a single compounding error can cause hundreds or even thousands of deaths. Yet before the 1940s, the law did not require drugs to be effective, only safe when used as directed. It was not until the 1940s that the Federal Trade Commission forced drug manufacturers to substantiate claims made about their products, at least those sold in interstate commerce. Instrumentation Although the 1920s and 1930s were especially fruitful for “soft” technologies such as antibiotics and vitamin production, these decades also produced several significant “hard” technologies—scientific instruments that transformed pharmaceutical R&D. Initially, one might think of the electron microscope, which was developed in 1931 in Germany. This early transmission electron microscope, invented by Max Knoll and Ernst Ruska, was essential for the future of pharmaceutical and biomedical research, but many other critical instruments came out of this era. Instrument production often brought people from disparate disciplines together on research teams, as physical chemists and physicists collaborated with biochemists and physiologists. New research fields were created along the boundary between chemistry and physics, and in the process, many new instruments were invented or adapted to molecular phenomena and biomolecular problems. One of the scientists who worked under the aegis of Warren Weaver, research administrator for the natural sciences at the Rockefeller Foundation, was Swedish chemist Theodor Svedberg. Svedberg’s early research was on colloids, and he began to develop high-speed centrifuges in the hope that they might provide an exact method for measuring the distribution of particle size in the solutions. In 1924, he developed the first ultracentrifuge, which generated a centrifugal force up to 5000 times the force of gravity. Later versions generated forces hundreds of thousands of times the force of gravity. Svedberg precisely determined the molecular weights of highly complex proteins, including hemoglobin. In later years, he performed studies in nuclear chemistry, contributed to the development of the cyclotron, and helped his student, Arne Tiselius, develop electrophoresis to separate and analyze proteins. Another essential instrument developed in this era was the pH meter with a glass electrode. Kenneth Goode first used a vacuum triode to measure pH in 1921, but this potentiometer was not coupled to a glass electrode until 1928, when two groups (at New York University and the University of Illinois) measured pH by using this combination. Rapid and inexpensive pH measurement was not a reality until 1934, however, when Arnold Beckman of the California Institute of Technology and corporate chemist Glen Joseph substituted a vacuum tube voltmeter for a galvanometer and assembled a sturdy measuring device with two vacuum tubes and a milliammeter. The portable pH meter was marketed in 1935 for $195. In the world of medical applications, the electrometer dosimeter was developed in the mid-1920s to assess exposure to ionizing radiation for medical treatment, radiation protection, and industrial exposure control. For clinical dosimetry and treatment planning, an ionization chamber connected to an electrometer was valued for its convenience, versatility, sensitivity, and reproducibility. It was not simply the invention of instruments, but the way research was organized around them, that made the 1920s and 1930s so fertile for biochemistry. Weaver was involved in encouraging and funding much of this activity, whether it was Svedberg’s work on molecular evolution or Linus Pauling’s use of X-ray diffraction to measure bond lengths and bond angles, and his development of the method of electron diffraction to measure the architecture of organic compounds. Inventing and developing new instruments allowed scientists to combine physics and chemistry and advance the field of pharmaceutical science. Radioisotopes Another powerful technological development that was refined during the 1920s and 1930s was the use of radioactive forms of elements—radioisotopes—in research. Hungarian chemist Georg von Hevesy introduced radioisotopes into experimental use in 1913, tracing the behavior of nonradioactive forms of selected elements; he later used a radioisotope of lead to trace the movement of lead from soil into bean plants. The radioactive tracer was an alternative to more arduous methods of measurement and study. By the late 1920s, researchers applied the tracer technique to humans by injecting dissolved radon into the bloodstream to measure the rate of blood circulation. Yet there were limits to the use of radioisotopes, owing to the fact that some important elements in living organisms do not possess naturally occurring radioisotopes. This difficulty was overcome in the early 1930s, when medical researchers realized that the cyclotron, or “atom smasher,” invented by physicist Ernest Lawrence, could be used to create radioisotopes for treatment and research. Radiosodium was first used in 1936 to treat several leukemia patients; the following year, Lawrence’s brother, John, used radiophosphorus to treat the same disease. A similar method was used to treat another blood disease, polycythemia vera, and soon it became a standard treatment for that malady. Joseph Hamilton and Robert Stone at the University of California, Berkeley, pioneered the use of cyclotron-produced radioisotopes for treating cancer in 1938; and one year later, Ernest Lawrence constructed an even larger atom smasher, known as the “medical cyclotron,” which would create additional radioisotopes in the hopes of treating cancer and other diseases. Thus began the age of “nuclear medicine,” in which the skills of physicists were necessary to produce materials critical to biochemical research. The new use of radioisotopes was a far cry from the quack medicines of the period that used the mystique of radioactivity to peddle radium pills and elixirs for human consumption—although in their ignorance, many legitimate doctors also did far more harm than good. The cost of producing radioisotopes was high, as cyclotrons often operated continuously at full power, requiring the attention of physicists around the clock. The human and financial resources of physics departments were strained in these early years, which made the contributions of foundations essential to the continuation of these innovative projects. Ultimately, as the century wore on, radioisotopes came into routine use, the federal government’s role increased immensely, and radioisotopes were mass-produced in the reactors of the Atomic Energy Commission. After World War II, nuclear medicine occupied a permanent place in pharmaceutical science. By the end of the 1930s, the work of scientists such as Svedberg, Tiselius, Banting, Dubos and Domagk, along with the vision of administrators such as Weaver, set the stage for the unparalleled developments of the antibiotic era to come. In addition, World War II, already beginning in Europe, spurred a wealth of research into perfecting known technologies and developing new ones—instruments and processes that would have a profound impact on the direction that biology and pharmacology would ultimately take. ANTIBIOTICS AND ISOTOPES (1940s) As the American public danced to the beat of the Big Band era, so did pharmacology swing into action with the upbeat tone of the dawning antibiotic era. The latter is the nickname most commonly used for the 1940s among scientists in the world of biotechnology and pharmaceuticals. The nickname is more than justified, given the numerous impressive molecules developed during the decade. Many “firsts” were accomplished in the drug discovery industry of the Forties, but no longer in a serendipitous fashion as before. Researchers were actually looking for drugs and finding them. To appreciate the events that paved the way for the advancements of the 1940s, one need only look back to 1939, when Rene Dubos of the Rockefeller Institute for Medical Research discovered and isolated an antibacterial compound—tyrothricin, from the soil microbe Bacillus brevis—capable of destroying Gram-positive bacteria.Before this discovery, penicillin and the sulfa drugs had been discovered “accidentally.” Dubos planned his experiment to search soil for microorganisms that could destroy organisms related to diseases. “It was a planned experiment. It wasn’t a chance observation,” explains H. Boyd Woodruff, who worked on the penicillin project at Merck in the early 1940s, “[and because the experiment] had been successful, it sort of opened the field in terms of looking at soil for microorganisms that kill disease organisms.” Fleming’s serendipity The most famous example of serendipity in the 20th century has to be the discovery of penicillin by Alexander Fleming as discussed in the previous chapter. Although Fleming observed the antibiotic properties of the mold Penicillium notatum in 1928, it was another 12 years before the active ingredient, penicillin, was isolated and refined. Of course, as in most of history, there are some inconsistencies. Fleming was not the first scientist to observe the antibacterial action of penicillin. In 1896 in Lyon, France, Ernest Augustin Duchesne studied the survival and growth of bacteria and molds, separately and together. He observed that the mold Penicillium glaucum had antibacterial properties against strains of both Escherichia coli and typhoid bacilli. The antibacterial properties of penicillin serendipitously surfaced at least three times during the course of scientific history before scientists used its power. And the third time was definitely a charm. Finding a magic bullet An Oxford University student of pathology, Howard Walter Florey’s early research interests involved mucus secretion and lysozyme—an antibacterial enzyme originally discovered by Fleming. The more he learned about the antibacterial properties of lysozyme and intestinal mucus, the more interested he became in understanding the actual chemistry behind the enzymatic reactions. However, he did not have the opportunity to work with chemists until 1935, when Florey hired Ernst Boris Chain to set up a biochemistry section in the department of pathology at the Sir William Dunn School of Pathology at Oxford. Because Chain was a chemist, Florey encouraged him to study the molecular action of lysozyme. Florey wanted to find out whether lysozyme played a role in duodenal ulcers and was less interested in its antibacterial properties. During a scientific literature search on bacteriolysis, Chain came upon Fleming’s published report of penicillin, which had, as Chain describes, “sunk into oblivion in the literature.” Chain thought that the active substance inducing staphylococcus lysis might be similar to lysozyme, and that their modes of action might also be similar. He set out to isolate penicillin to satisfy his own scientific curiosity and to answer a biological problem—what reaction lysozyme catalyzes—not to find a drug. The scientific collaborations and discussions between Chain and Florey eventually laid the foundation for their 1939 funding application to study the antimicrobial products of microorganisms. It never crossed their minds that one of these antimicrobial products would be the next magic bullet. The timing of their funded research is also significant—it occurred within months of Great Britain’s declaration of war with Germany and the beginning of World War II. Because of its activity against staphylococcus, Fleming’s penicillin was one of the first compounds chosen for the study. The first step toward isolating penicillin came in March 1940 at the suggestion of colleague Norman G. Heatley. The team extracted the acidified culture filtrate into organic solution and then re-extracted penicillin into a neutral aqueous solution. In May, Florey examined the chemotherapeutic effects of penicillin by treating four of eight mice infected with Streptomyces pyogenes. The mice treated with penicillin survived, whereas the other 4 died within 15 hours. In September, Henry Dawson and colleagues confirmed the antibiotic properties of penicillin by taking a bold step and injecting penicillin into a patient at Columbia Presbyterian Hospital (New York). With the help of Chain, Heatley, Edward P. Abraham, and other Dunn School chemists, Florey was able to scrape up enough penicillin to perform clinical trials at the Radcliffe Infirmary in Oxford in February 1941. The first patient treated was dying of S. aureus and S. pyogenes. Treatment with penicillin resulted in an amazing recovery, but because of insufficient quantities of the drug, the patient eventually died after a relapse. Over the next three months, five other patients responded well when treated with penicillin. All of these patients were seriously ill with staphylococcal or streptococcal infections that could not be treated with sulfonamide. These trials proved the effectiveness of penicillin when compared to the sulfa drugs, which at the time were considered the gold standard for treating infections. Producing penicillin Florey had difficulties isolating the quantities of penicillin required to prove its value. In the early years, the Oxford team grew the mold by surface culture in anything they could lay their hands on. Because of the war, they couldn’t get the glass flasks they wanted, so they used bedpans until Florey convinced a manufacturer to make porcelain pots, which incidentally resembled bedpans. Britain was deep into the war, and the British pharmaceutical industry did not have the personnel, material, or funds to help Florey produce penicillin. Florey and Heatley came to the United States in June 1941 to seek assistance from the American pharmaceutical industry. They traveled around the country but could not garner interest for the project. Because of the as yet ill-defined growing conditions for P. notatum and the instability of the active compound, the yield of penicillin was low and it was not economically feasible to produce. Florey and Heatley ended up working with the U.S. Department of Agriculture’s Northern Regional Research Laboratory in Peoria, IL. The agricultural research center had excellent fermentation facilities, but more importantly—unlike any other facility in the country—it used corn steep liquor in the medium when faced with problematic cultures. This liquor yielded remarkable results for the penicillin culture. The production of penicillin increased by more than 10-fold, and the resulting penicillin was stable. It turns out that the penicillin (penicillin G) produced at the Peoria site was an entirely different compound from the penicillin (penicillin F) produced in Britain. Fortunately for all parties involved, penicillin G demonstrated the same antibacterial properties against infections as penicillin F. With these new developments, Merck, Pfizer, and Squibb agreed to collaborate on the development of penicillin. By this time, the United States had entered the war, and the U.S. government was encouraging pharmaceutical companies to collaborate and successfully produce enough penicillin to treat war-related injuries. By 1943, several U.S. pharmaceutical companies were mass-producing purified penicillin G (~21 billion dosage units per month), and it became readily available to treat bacterial infections contracted by soldiers. In fact, by 1944, there was sufficient penicillin to treat all of the severe battle wounds incurred on D-day at Normandy. Also, diseases like syphilis and gonorrhea could suddenly be treated more easily than with earlier treatments, which included urethra cleaning and doses of noxious chemicals such as mercury or Salvarsan. The Americans continued to produce penicillin at a phenomenal rate, reaching nearly 7 trillion units per month in 1945. Fleming, Florey, and Chain were recognized “for the discovery of penicillin and its curative effect in various infectious diseases” in 1945 when they received the Nobel Prize in Physiology or Medicine. But all magic bullets lose their luster, and penicillin was no different. Dubos had the foresight to understand the unfortunate potential of antibiotic-resistant bacteria and encouraged prudent use of antibiotics. As a result of this fear, Dubos stopped searching for naturally occurring compounds with antibacterial properties. As early as 1940, Abraham and Chain identified a strain of S. aureus that could not be treated with penicillin. This seemingly small, almost insignificant event foreshadowed the wave of antibiotic-resistant microorganisms that became such a problem throughout the medical field toward the end of the century. Malaria and quinine Although penicillin was valuable against the battle-wound infections and venereal diseases that have always afflicted soldiers, it was not effective against the malaria that was killing off the troops in the mosquito-ridden South Pacific. The Americans entered Guadal canal in June 1942, and by August there were 900 cases of malaria; in September, there were 1724, and in October, 2630. By December 1942, more than 8500 U.S. soldiers were hospitalized with malaria. Ninety percent of the men had contracted the disease, and in one hospital, as many as eight of every 10 soldiers had malaria rather than combat-related injuries. The only available treatment, however, was the justifiably unpopular drug Atabrine. Besides tasting bitter, the yellow Atabrine pills caused headaches, nausea, vomiting, and in some cases, temporary psychosis. It also seemed to leave a sickly hue to the skin and was falsely rumored to cause impotence. Nevertheless, it was effective and saved lives. Firms such as Abbott, Lilly, Merck, and Frederick Stearns assured a steady supply of Atabrine, producing 3.5 billion tablets in 1944 alone. But Atabrine lacked the efficacy of quinine, which is isolated from cinchona, an evergreen tree native to the mountains of South and Central America. Unfortunately, the United States did not have a sufficient supply of quinine in reserve when the war broke out. As a result, the U.S. government established the Cinchona Mission in 1942. Teams of botanists, foresters, and assistants went to South America to find and collect quinine-rich strains of the tree—a costly, strenuous, and time-consuming task. Out of desperation, research to develop antimalarials intensified. As an unfortunate example of this desperation, prison doctors in the Chicago area experimentally infected nearly 400 inmates with malaria during their search for a therapeutic. Although aware that they were helping the war effort, the prisoners were not given sufficient information about the details and risks of the clinical experiments. After the war, Nazi doctors on trial for war crimes in Nuremberg referred to this incident as part of their defense for their criminal treatment of prisoners while aiding the German war effort. In 1944, William E. Doering and Robert B. Woodward synthesized quinine—a complex molecular structure—from coal tar. Woodward’s achievements in the art of organic synthesis earned him the 1965 Nobel Prize in Chemistry. Chloroquine, another important antimalarial, was synthesized and studied under the name of Resochin by the German company Bayer in 1934 and rediscovered in the mid-1940s. Even though chloroquine-resistant parasites cause illness throughout the world, the drug is still the primary treatment for malaria. Streptomycin and tuberculosis When Dubos presented his results with tyrothricin at the Third International Congress for Microbiology in New York in 1939, Selman A. Waksman was there to see it. The successful development of penicillin and the discovery of tyrothricin made Waksman realize the enormous potential of soil as a source of druglike compounds. He immediately decided to focus on the medicinal uses of antibacterial soil microbes. In 1940, Woodruff and Waksman isolated and purified actinomycin from Actinomyces griseus (later named Streptomyces griseus), which led to the discovery of many other antibiotics from that same group of microorganisms. Actinomycin attacks Gram-negative bacteria responsible for diseases like typhoid, dysentery, cholera, and undulant fever and was the first antibiotic purified from an actinomycete. Considered too toxic for the treatment of diseases in animals or humans, actinomycin is primarily used as an investigative tool in cell biology. In 1942, the two researchers isolated and purified streptothricin, which prevents the proliferation of Mycobacterium tuberculosis but is also too toxic for human use. A couple of years later, in 1944, Waksman, with Albert Schatz and Elizabeth Bugie, isolated the first aminoglycoside, streptomycin, from S. griseus. Like penicillin, aminoglycosides decrease protein synthesis in bacterial cells, except that streptomycin targets Gram-positive organisms instead of Gram-negatives. Waksman studied the value of streptomycin in treating bacterial infections, especially tuberculosis. In 1942, several hundred thousand deaths resulted from tuberculosis in Europe, and another 5 to 10 million people suffered from the disease. Although sulfa drugs and penicillin were readily available, they literally had no effect. Merck immediately started manufacturing streptomycin with the help of Woodruff. A consultant for Merck, Waksman sent Woodruff to Merck to help with the penicillin project, and after finishing his thesis, Woodruff continued working there. Simultaneously, studies by W. H. Feldman and H. C. Hinshaw at the Mayo Clinic confirmed streptomycin’s efficacy and relatively low toxicity against tuberculosis in guinea pigs. On November 20, 1944, doctors administered streptomycin for the first time to a seriously ill tuberculosis patient and observed a rapid, impressive recovery. No longer unconquerable, tuberculosis could be tamed and beaten into retreat. In 1952, Waksman was awarded the Nobel Prize in Physiology or Medicine for his discovery of streptomycin—1 of 18 antibiotics discovered under his guidance—and its therapeutic effects in patients suffering from tuberculosis. Merck had just developed streptomycin and moved it into the marketplace when the company stumbled upon another great discovery. At the time, doctors treated patients with pernicious anemia by injecting them with liver extracts, which contained a factor required for curing and controlling the disease. When patients stopped receiving injections, the disease redeveloped. The Merck chemists had been working on isolating what was called the pernicious anemia factor from liver extracts, and they decided to look at the cultures grown by Woodruff and other microbiologists at Merck, to see if one of the cultures might produce the pernicious anemia factor. They found a strain of S. griseus similar to the streptomycin-producing strain that made the pernicious anemia factor. With the help of Mary Shorb’s Lactobacillus lactis assay to guide the purification and crystallization of the factor, Merck scientists were able to manufacture and market the factor as a cure for pernicious anemia. The factor turned out to be a vitamin, and it was later named vitamin B12. As Woodruff describes the period, “So, we jumped from penicillin to streptomycin to vitamin B12. We got them 1-2-3, bang-bang-bang.” Merck struck gold three times in a row. The United Kingdom’s only woman Nobel laureate, Dorothy Crowfoot Hodgkin, solved the molecular structure of vitamin B12 in 1956—just as she had for penicillin in the 1940s, a discovery that was withheld from publication until World War II was over. The continuing search After developing penicillin, U.S. pharmaceutical companies continued to search for “antibiotics,” a term coined by P. Vuillemin in 1889 but later defined by Waksman in 1947 as those chemical substances “produced by microbes that inhibit the growth of and even destroy other microbes.” In 1948, Benjamin M. Duggar, a professor at the University of Wisconsin and a consultant to Lederle, isolated chlortetracycline from Streptomyces aureofaciens. Chlortetracycline, also called aureomycin, was the first tetracycline antibiotic and the first broad-spectrum antibiotic. Active against an estimated 50 disease organisms, aureomycin works by inhibiting protein synthesis. The discovery of the tetracycline ring system also enabled further development of other important antibiotics. Other antibiotics with inhibitory effects on cell wall synthesis were also discovered in the 1940s and include cephalosporin and bacitracin. Another ß-lactam antibiotic, cephalosporin was first isolated from Cephalosporium acremonium in 1948 by Guiseppe Brotzu at the University of Cagliari in Italy. Bacitracin, first derived from a strain of Bacillus subtilis, is active against Gram-positive bacteria and is used topically to treat skin infections. Nonantibiotic therapeutics Even though Lederle Laboratories was a blood processing plant during World War II, it evolved into a manufacturer of vitamins and nutritional products, including folic acid. Sidney Farber, a cancer scientist at Boston’s Children’s Hospital, was testing the effects of folic acid on cancer. Some of his results, which now look dubious, suggested that folic acid worsened cancer conditions, inspiring chemists at Lederle to make antimetabolites—structural mimics of essential metabolites that interfere with any biosynthetic reaction involving the intermediates—resembling folic acid to block its action. These events led to the 1948 development of methotrexate, one of the earliest anticancer agents and the mainstay of leukemia chemotherapy. But the pioneer of designing and synthesizing antimetabolites that could destroy cancer cells was George Hitchings, head of the department of biochemistry at Burroughs Wellcome Co. In 1942, Hitchings initiated his DNA-based antimetabolite program, and in 1948, he and Gertrude Elion synthesized and demonstrated the anticancer activity of 2,6-diaminopurine. By fine-tuning the structure of the toxic compound, Elion synthesized 6-mercaptopurine, a successful therapeutic for treating acute leukemia. Hitchings, Elion, and Sir James W. Black won the Nobel Prize in Physiology or Medicine in 1988 for their discoveries of “important principles for drug treatment,” which constituted the groundwork for rational drug design. The discovery of corticosteroids as a therapeutic can be linked to Thomas Addison, who made the connection between the adrenal glands and the rare Addison’s disease in 1855. But it wasn’t until Edward Calvin Kendall at the Mayo Clinic and Thadeus Reichstein at the University of Basel independently isolated several hormones from the adrenal cortex that corticosteroids were used to treat a more widespread malady. In 1948, Kendall and Philip S. Hench demonstrated the successful treatment of patients with rheumatoid arthritis using cortisone. Kendall, Reichstein, and Hench received the 1950 Nobel Prize in Physiology or Medicine for determining the structure and biological effects of adrenal cortex hormones. One of the first therapeutic drugs to prevent cardiovascular disease also came from this period. While investigating the mysterious deaths of farm cows, Karl Paul Link at the University of Wisconsin proved that the loss of clotting ability in cattle was linked to the intake of sweet clover. He and his colleagues then isolated the anticoagulant and blood thinner dicoumarol (warfarin) from coumarin, a substance found in sweet clover, in 1940. Many other advances The synthesis, isolation, and therapeutic applications of miracle drugs may be the most well-remembered discoveries of the 1940s for medical chemists and biochemists, but advances in experimental genetics, biology, and virology were also happening. These advances include isolating the influenza B virus in 1940, by Thomas Francis at New York University and, independently, by Thomas Pleines Magill. Also in 1940, at the New York Hospital–Cornell University Medical Center, Mary Loveless succeeded in blocking the generation of immunotherapy-induced antibodies using pollen extracts. Routine use of the electron microscope in virology followed the first photos of tobacco-mosaic virus by Helmut Ruska, an intern at the Charité Medical School of Berlin University, in 1939; and the 1940s also saw numerous breakthroughs in immunology, including the first description of phagocytosis by a neutrophil. In 1926, Hermann J. Muller, a professor at the University of Texas at Austin, reported the identification of several irradiation-induced genetic alterations, or mutations, in Drosophila that resulted in readily observed traits. This work, which earned Muller the Nobel Prize in Physiology or Medicine in 1946, enabled scientists to recognize mutations in genes as the cause of specific phenotypes, but it was still unclear how mutated genes led to the observed phenotypes. In 1935, George Wells Beadle began studying the development of eye pigment in Drosophila with Boris Ephrussi at the Institut de Biologie Physico-Chimique in Paris. Beadle then collaborated with Edward Lawrie Tatum when they both joined Stanford in 1937—Beadle as a professor of biology (genetics) and Tatum as a research associate in the department of biological sciences. Tatum, who had a background in chemistry and biochemistry, handled the chemical aspects of the Drosophila eye-color study. Beadle and Tatum eventually switched to the fungus Neurospora crassa, a bread mold. After producing mutants of Neurospora by irradiation and searching for interesting phenotypes, they found several auxotrophs—strains that grow normally on rich media but cannot grow on minimal medium. Each mutant required its own specific nutritional supplement, and each requirement correlated to the loss of a compound normally synthesized by the organism. By determining that each mutant evoked a deficiency in a specific metabolic pathway, which was known to be controlled by enzymes, Beadle and Tatum concluded in a 1940 report that each gene produced a single enzyme, also called the “single gene–single enzyme” concept. The two scientists shared the Nobel Prize in Physiology or Medicine in 1958 for discovering that genes regulate the function of enzymes and that each gene controls a specific enzyme. Also recognized with the same prize in 1958 was Joshua Lederberg. As a graduate student in Tatum’s laboratory in 1946, Lederberg found that some plasmids enable bacteria to transfer genetic material to each other by forming direct cell–cell contact in a process called conjugation. He also showed that F (fertility) factors allowed conjugation to occur. In addition, Lederberg defined the concepts of generalized and specialized transduction, collaborated with other scientists to develop the selection theory of antibody formation, and demonstrated that penicillin-susceptible bacteria could be grown in the antibiotic’s presence if a hypotonic medium was used. In the field of virology, John Franklin Enders, Thomas H. Weller, and Frederick Chapman Robbins at the Children’s Hospital Medical Center in Boston figured out in 1949 how to grow poliovirus in test-tube cultures of human tissues—a technique enabling the isolation and study of viruses. Polio, often referred to as infantile paralysis, was one of the most feared diseases of the era. These researchers received the Nobel Prize in Physiology or Medicine in 1954. Salvador Luria, at Indiana University, and Alfred Day Hershey, at Washington University’s School of Medicine, demonstrated that the mutation of bacteriophages makes it difficult for a host to develop immunity against viruses. In 1942, Thomas Anderson and Luria photographed and characterized E. coli T2 bacteriophages using an electron microscope. Luria and Max Delbrück, at Vanderbilt University, used statistical methods to demonstrate that inheritance in bacteria follows Darwinian principles. Luria, Hershey, and Delbrück were awarded the Nobel Prize in Physiology or Medicine in 1969 for elucidating the replication mechanism and genetic structure of viruses. Although these discoveries were made outside of the pharmaceutical industry, their applications contributed enormously to understanding the mechanisms of diseases and therapeutic drugs. Biological and chemical warfare Biological warfare—the use of disease to harm or kill an adversary’s military forces, population, food, and livestock—can involve any living microorganism, nonliving virus, or bioactive substance deliverable by conventional artillery. The history of biological warfare can be traced back to the Romans, who used dead animals to infect their enemies’ water supply. The United States started a biological warfare program in 1942 after obtaining Japanese data about the destructive use of chemical and biological agents from pardoned war criminals. Japan sprayed bubonic plague over parts of mainland China on five occasions in 1941. Despite the fact that the spraying was ineffective, the attempts prompted the United States to develop its biological warfare program. Later, the developing Cold War further stimulated this research in the United States—and in the Soviet Union. Ironically, the first chemotherapeutic agent for cancer came from an early instance of chemical warfare. Initially used as a weapon in World War I, mustard gas proved useful in treating mice and a person with lymphoma in 1942, when Alfred Gilman and Fred Phillips experimentally administered the chemical weapon as a therapeutic. Because the patient showed some improvement, chemical derivatives of mustard gas were developed and used to treat various cancers. The Nuremberg Code Not only did World War II encourage the discovery and development of antibiotics and antidisease drugs, it also instigated the need to define what constitutes permissible medical experiments on human subjects. The Nazis performed cruel and criminal “medical” experiments on Jews and other prisoners during the war. In 1949, the Nuremberg Code was established in an effort to prevent medical crimes against humanity. The Code requires that individuals enrolled in clinical trials give voluntary consent. The experiment must hypothetically achieve useful results for the good of society, be performed by scientifically qualified persons, and be derived from experiments on animal models that suggest the anticipated outcome will justify human clinical experiments. The code also emphasizes that all physical and mental suffering must be avoided and that precautions must be taken to protect the human subject if injury or disability results from the experiment. In achieving its goals, the Nuremberg Code necessarily empowers the human subject and holds the researcher responsible for inflicting unnecessary pain and suffering on the human subject. On the practical level, it was not until the 1960s that institutionalized protections for subjects in clinical trials and human experimentation were put into place. “Swing” time The 1940s ended with the antibiotic era in full swing and with a host of wartime advancements in fermentation and purification technologies changing the drug development process. Penicillin and DDT became the chemical markers of the age, promising to heal the world—curing the plagues and killing the plague carriers. The radioisotopes now easily produced through advances in technology promoted by the war were becoming routinely available for health research, as the era of computer-aided drug analysis began. The baby boom launched by postwar U.S. prosperity produced the first generation born with the expectation of health through drugs and medical intervention. Because of these new possibilities, health became a political as well as a social issue. The leading role science played in the Allied victory gave way in the postwar 1940s to its new role as medical savior. The new technologies that became available in the 1940s—including partition chromatography, infrared and mass spectrometry, as well as nuclear magnetic resonance (NMR)—would eventually become critical to pharmaceutical progress. Prescriptions & polio (1950s) The 1950s began amid a continuing wave of international paranoia. The Cold War intensified in the late 1940s as the West responded to the ideological loss of China to communism and the very real loss of atom bomb exclusivity to the Soviets. The first few years of the 1950s heated up again with the Korean conflict. On the home front, World War II– related science that was now declassified, together with America’s factories now turned to peace, shaped a postwar economic boom. Driven by an unprecedented baby boom, the era’s mass consumerism focused on housing, appliances, automobiles, and luxury goods. Technologies applied to civilian life included silicone products, microwave ovens, radar, plastics, nylon stockings, long-playing vinyl records, and computing devices. New medicines abounded, driven by new research possibilities and the momentum of the previous decade’s “Antibiotic Era.” A wave of government spending was spawned by two seminal influences: a comprehensive new federal science and technology policy and the anti-“Red” sentiment that dollars spent for science were dollars spent for democracy. While improved mechanization streamlined production in drug factories, the DNA era dawned. James Watson and Francis Crickdetermined the structure of the genetic material in 1953. Prescription and nonprescription drugs were legally distinguished from one another for the first time in the United States as the pharmaceutical industry matured. Human cell culture and radioimmunoassays developed as key research technologies; protein sequencing and synthesis burgeoned, promising the development of protein drugs. In part because of Cold War politics, in part because the world was becoming a smaller place, global health issues took center stage. Fast foods and food additives became commonplace in theWest. “The Pill” was developed and first tested in Puerto Rico. Ultrasound was adapted to fetal monitoring. Gas chromatography (GC), mass spectrometry, and polyacrylamide gel electrophoresis began transforming drug research, as did the growth of the National Institutes of Health (NIHI) and the National Science Foundation (NSF). The foundations of modern immunology were laid as the pharmaceutical industry moved ever forward in mass-marketing through radio and the still-novel format of television. But above all, through the lens of this time in the Western world, was the heroic-scientist image of Jonas Salk, savior of children through the conquest of polio via vaccines. Antibiotics redux The phenomenal success of antibiotics in the 1940s spurred the continuing pursuit of more and better antibacterial compounds from a variety of sources, especially from natural products in soils and synthetic modifications of compounds discovered earlier. In 1950, the antibiotic Nystatin was isolated from Streptomyces noursei obtained from soil in Virginia. In 1952, erythromycin was first isolated from S. erythreus from soil in the Philippines. Other antibiotics included Novobiocin (1955), isolated from S. spheroides from Vermont; Vancomycin (1956) from S. orientalis from soils in Borneo and Indiana; and Kanamycin (1957) from S. kanamyceticus from Japan. The search for antibiotic compounds also led researchers in Great Britain to discover, in 1957, an animal glycoprotein (interferon) with antiviral activity. Not only did the development of antibiotics that began in the 1940s lead to the control of bacterial infections, it also permitted remarkable breakthroughs in the growth of tissue culture in the 1950s. These breakthroughs enabled the growth of polio and other viruses in animal cell cultures rather than in whole animals, and permitted a host of sophisticated physiological studies that had never before been possible. Scientists were familiar with the concepts of tissue culture since the first decades of the century, but routine application was still too difficult. After antibiotics were discovered, such research no longer required an “artist in biological technique” to maintain the requisite sterile conditions in the isolation, maintenance, and use of animal cells, according to virus researcher Kingsley F. Sanders. In 1957, he noted that the use of antibiotics in tissue culture made the process so easy that “even an amateur in his kitchen can do it.” Funding medicine In the 1950s, general science research, particularly biological research, expanded in the United States to a great extent because of the influence of Vannevar Bush, presidential science advisor during and immediately after World War II. His model, presented in the 1945 report to the President, Science: The Endless Frontier, set the stage for the next 50 years of science funding. Bush articulated a linear model of science in which basic research leads to applied uses. He insisted that science and government should continue the partnership forged in the 1940s with the Manhattan Project (in which he was a key participant) and other war-related research. As early as 1946, Bush argued for creating a national science funding body. The heating up of the Cold War, as much as anything else, precipitated the 1950 implementation of his idea in the form of the National Science Foundation—a major funnel for government funding of basic research, primarily for the university sector. It was a federal version of the phenomenally successful Rockefeller Foundation. The new foundation had a Division of Biological and Medical Sciences, but its mission was limited to supporting basic research so that it wouldn’t compete with the more clinically oriented research of the NIH. The NIH rode high throughout the 1950s, with Congress regularly adding $8 million to $15 million to the NIH budget proposed by the first Eisenhower administration. By 1956, the NIH budget had risen to almost $100 million. By the end of the decade, the NIH was supporting some 10,000 research projects at 200 universities and medical schools at a cost of $250 million. Other areas of government also expanded basic medical research under the Bush vision. In 1950, for example, the Atomic Energy Commission received a $5 million allocation from Congress specifically to relate atomic research to cancer treatment. In this same vein, in 1956, Oak Ridge National Laboratory established a medical instruments group to help promote the development of technology for disease diagnostics and treatment that would lead, in conjunction with advances in radioisotope technology, to a new era of physiologically driven medicine. Part of this research funding was under the auspices of the Atoms for Peace program and led to the proliferation of human experiments using radioactive isotopes, often in a manner that would horrify a later generation of Americans with its cavalier disregard for participants’ rights. Science funding, including medical research, received an additional boost with the 1957 launch of the first orbital satellite. The Soviet sputnik capped the era with a wave of science and technology fervor in industry, government, and even the public schools. The perceived “science gap” between the United States and the Soviet Union led to the 1958 National Defense Education Act. The act continued the momentum of government-led education that started with the GI Bill to provide a new, highly trained, and competent workforce that would transform industry. This focus on the importance of technology fostered increased reliance on mechanized mass-production techniques. During World War II, the pharmaceutical industry had learned its lesson—that bigger was better in manufacturing methods—as it responded to the high demand for penicillin. Private funds were also increasingly available throughout the 1950s—and not just from philanthropic institutions. The American Cancer Society and the National Foundation for Infantile Paralysis were two of the largest public disease advocacy groups that collected money from the general public and directed the significant funds to scientific research. This link between the public and scientific research created, in some small fashion, a sense of investment in curing disease, just as investing in savings bonds and stamps in the previous decade had created a sense of helping to win World War II. The war on polio Perhaps the most meaningful medical story to people of the time was that of Jonas Salk and his “conquest of polio.” Salk biographer Richard Carter described the response to the April 12, 1955, vaccine announcement: “More than a scientific achievement, the vaccine was a folk victory, an occasion for pride and jubilation…. People observed moments of silence, rang bells, honked horns, … closed their schools or convoked fervid assemblies therein, drank toasts, hugged children, attended church.” The public felt a strong tie to Salk’s research in part because he was funded by the National Foundation for Infantile Paralysis. Since 1938, the organization’s March of Dimes had collected small change from the general public to fund polio research. The group managed to raise more money than was collected for heart disease or even cancer. The Salk vaccine relied on the new technology of growing viruses in cell cultures, specifically in monkey kidney cells (first available in 1949). Later, the human HeLa cell line was used as well. The techniques were developed by John F. Enders (Harvard Medical School), Thomas H. Weller (Children’s Medical Center, Boston), and Frederick C. Robbins (Case Western Reserve University, Cleveland), who received the 1954 Nobel Prize in Physiology or Medicine for their achievement. Salk began preliminary testing of his polio vaccine in 1952, with a massive field trial in the United States in 1954. According to Richard Carter, a May 1954 Gallup Poll found that “more Americans were aware of the polio field trial than knew the full name of the President of the United States.” Salk’s vaccine was a killed-virus vaccine that was capable of causing the disease only when mismanaged in production. This unfortunately happened within two weeks of the vaccine’s release. The CDC Poliomyelitis Surveillance Unit was immediately established; the popularly known “disease detectives” traced down the problem almost immediately, and the “guilty” vaccine lot was withdrawn. It turned out that Cutter Laboratories had released a batch of vaccine with live contaminants that tragically resulted in at least 260 cases of vaccine-induced polio. Ironically, these safety problems helped promote an alternative vaccine that used live rather than killed virus. In 1957, the attenuated oral polio vaccine was finally developed by Albert Sabin and became the basis for mass vaccinations in the 1960s. The live vaccine can infect a small percentage of those inoculated against the disease with active polio, primarily those with compromised immune systems, and is also a danger to immunocompromised individuals who have early contact with the feces of vaccinated individuals. In its favor, it provides longer lasting immunity and protection against gastrointestinal reinfection, eliminating the reservoir of polio in the population. The debate that raged in the 1950s over Salk versus Sabin (fueled at the time by a history of scientific disputes between the two men) continues today: Some countries primarily use the injected vaccine, others use the oral, and still others use one or the other, depending on the individual patient. Instruments and assays The 1950s saw a wave of new instrumentation, some of which, although not originally used for medical purposes, was eventually used in the medical field. In 1951, image-analyzing microscopy began. By 1952, thin sectioning and fixation methods were being perfected for electron microscopy of intracellular structures, especially mitochondria. In 1953, the first successful open-heart surgery was performed using the heart–lung machine developed by John H. Gibbon Jr. in Philadelphia. Of particular value to medical microbiology and ultimately to the development of biotechnology was the production of an automated bacterial colony counter. This type of research was first commissioned by the U.S. Army Chemical Corps. Then the Office of Naval Research and the NIH gave a significant grant for the development of the Coulter counter, commercially introduced as the Model A. A.J.P. Martin in Britain developed gas–liquid partition chromatography in 1952. The first commercial devices became available three years later, providing a powerful new technology for chemical analysis. In 1954, Texas Instruments introduced silicon transistors—a technology encompassing everything from transistorized analytical instruments to improved computers and, for the mass market, miniaturized radios. The principle for electromagnetic microbalances was developed near the middle of the decade, and a prototype CT scanner was unveiled. In 1958, amniocentesis was developed, and Scottish physician Ian McDonald pioneered the use of ultrasound for diagnostics and therapeutics. Radiometer micro pH electrodes were developed by Danish chemists for bedside blood analysis. In a further improvement in computing technology, Jack Kilby at Texas Instruments developed the integrated circuit in 1958. In 1959, the critical technique of polyacrylamide gel electrophoresis (PAGE) was in place, making much of the coming biotechnological analysis of nucleic acids and proteins feasible. Strides in the use of atomic energy continued apace with heavy government funding. In 1951, Brookhaven National Laboratory opened its first hospital devoted to nuclear medicine, followed seven years later by a Medical Research Center dedicated to the quest for new technologies and instruments. By 1959, the Brookhaven Medical Research Reactor was inaugurated, making medical isotopes significantly cheaper and more available for a variety of research and therapeutic purposes. In one of the most significant breakthroughs in using isotopes for research purposes, in 1952, Rosalyn Sussman Yalow, working at the Veterans Hospital in the Bronx in association with Solomon A. Berson, developed the radioimmunoassay (RIA) for detecting and following antibodies and other proteins and hormones in the body. Physiology explodes The development of new instruments, radioistopes, and assay techniques found rapid application in the realm of medicine as research into general physiology and therapeutics prospered. In 1950, for example, Konrad Emil Bloch at Harvard University used carbon-13 and carbon-14 as tracers in cholesterol buildup in the body. Also in 1950, Albert Claude of the Université Catolique de Louvain in Belgium discovered the endoplasmic reticulum using electron microscopy. That same year, influenza type C was discovered. New compounds and structures were identified in the human body throughout the decade. In 1950, GABA (gamma-aminobutyric acid ) was identified in the brain. Soon after that, Italian biologist Rita Levi-Montalcini demonstrated the existence of a nerve growth hormone. In Germany, F.F.K. Lynen isolated the critical enzyme cofactor, acetyl-CoA, in 1955. Human growth hormone was isolated for the first time in 1956. That same year, William C. Boyd of the Boston University Medical School identified 13 “races” of humans based on blood groups. Breakthroughs were made that ultimately found their way into the development of biotechnology. By 1952, Robert Briggs and Thomas King, developmental biologists at the Institute for Cancer Research in Philadelphia, successfully transplanted frog nuclei from one egg to another—the ultimate forerunner of modern cloning techniques. Of tremendous significance to the concepts of gene therapy and specific drug targeting, sickle cell anemia was shown to be caused by one amino acid difference between normal and sickle hemoglobin (1956–1958). Although from today’s perspective it seems to have occurred surprisingly late, in 1956 the human chromosome number was finally revised from the 1898 estimate of 24 pairs to the correct 23 pairs. By 1959, examination of chromosome abnormalities in shape and number had become an important diagnostic technique. That year, it was determined that Down’s syndrome patients had 47 chromosomes instead of 46. As a forerunner of the rapid development of immunological sciences, in 1959 Australian virologist Frank Macfarlane Burnet proposed his clonal selection theory of antibody production, which stated that antibodies were selected and amplified from preexisting rather than instructionally designed templates. A rash of new drugs New knowledge (and, as always, occasional serendipity) led to new drugs. The decade began with the Mayo Clinic’s discovery of cortisone in 1950—a tremendous boon to the treatment of arthritis. But more importantly, it saw the first effective remedy for tuberculosis. In 1950, British physician Austin Bradford Hill demonstrated that a combination of streptomycin and para-aminosalicylic acid (PAS) could cure the disease, although the toxicity of streptomycin was still a problem. By 1951, an even more potent antituberculosis drug was developed simultaneously and independently by the Squibb Co. and Hoffmann-LaRoche. Purportedly after the death of more than 50,000 mice (part of a new rapid screening method developed by Squibb to replace the proverbial guinea pigs as test animals) and the examination of more than 5000 compounds, isonicotinic acid hydrazide proved able to protect against a lethal dose of tubercle bacteria. It was marketed ultimately as isoniazid and proved especially effective in mixed dosage with streptomycin or pas. In 1951, monoamine oxidase (MAO) inhibitors were introduced to treat psychosis. In 1952, reserpine was isolated from rauwolfia and eventually was used for treating essential hypertension. But in 1953, the rauwolfia alkaloid was used as the first of the tranquilizer drugs. The source plant came from India, where it had long been used as a folk medicine. The thiazide drugs were also developed in this period as diuretics for treating high blood pressure. In 1956, halothane was introduced as a general anesthetic. In 1954, the highly touted chlorpromazine (Thorazine) was approved as an antipsychotic in the United States. It had started as an allergy drug developed by the French chemical firm Rhône-Poulenc, and it was noticed to have “slowed down” bodily processes. Also in 1954, the FDA approved BHA (butylated hydroxyanisole) as a food preservative; coincidentally, McDonald’s was franchised that same year. It soon became the largest “fast food” chain. Although not really a new “drug” (despite numerous fast food “addicts”), the arrival and popularity of national fast food chains (and ready-made meals such as the new TV dinners in the supermarket) were the beginning of a massive change in public nutrition and thus, public health. Perhaps the most dramatic change in the popularization of drugs came with the 1955 marketing of meprobamate (first developed by Czech scientist Frank A. Berger) as Miltown (by Wallace Laboratories) and Equanil (by Wyeth). This was the first of the major tranquilizers or anti-anxiety compounds that set the stage for the 1960s “drug era.” The drug was so popular that it became iconic in American life. (The most popular TV comedian of the time once referred to himself as “Miltown” Berle.) Unfortunately, meprobamate also proved addictive. In 1957, British researcher Alick Isaacs and J. Lindenman of the National Institute for Medical Research, Mill Hill, London, discovered interferon—a naturally occurring antiviral protein, although not until the 1970s (with the advent of gene-cloning technology) would it become routinely available for drug use. In 1958, a saccharin-based artificial sweetener was introduced to the American public. That year also marked the beginning of the thalidomide tragedy (in which the use of a new tranquilizer in pregnant women caused severe birth defects), although it would not become apparent until the 1960s. In 1959, Haldol (haloperidol) was first synthesized for treating psychotic disorders. Blood products also became important therapeutics in this decade, in large part because of the 1950 development of methods for fractionating blood plasma by Edwin J. Cohn and colleagues. This allowed the production of numerous blood-based drugs, including fraction X (1956), a protein common to both the intrinsic and extrinsic pathways of blood clotting, and fraction VIII (1957), a blood-clotting protein used for treating hemophilia. The birth of birth control Perhaps no contribution of chemistry in the second half of the 20th century had a greater impact on social customs than the development of oral contraceptives. Several people were important in its development—among them Margaret Sanger, Katherine McCormick, Russell Marker, Gregory Pincus, and Carl Djerassi. Sanger was a trained nurse who was a supporter of radical, left-wing causes. McCormick was the daughter-in-law of Cyrus McCormick, founder of International Harvester, whose fortune she inherited when her husband died. Both were determined advocates of birth control as the means to solving the world’s overpopulation. Pincus (who founded the Worcester Foundation for Experimental Biology) was a physiologist whose research interests focused on the sexual physiology of rabbits. He managed to fertilize rabbit eggs in a test tube and got the resulting embryos to grow for a short time. The feat earned him considerable notoriety, and he continued to gain a reputation for his work in mammalian reproductive biology. Sanger and McCormick approached Pincus and asked him to produce a physiological contraceptive. He agreed to the challenge, and McCormick agreed to fund the project. Pincus was certain that the key was the use of a female sex hormone such as progesterone. It was known that progesterone prevented ovulation and thus was a pregnancy-preventing hormone. The problem was finding suitable, inexpensive sources of the scarce compound to do the necessary research. Enter American chemist Russell Marker. Marker’s research centered on converting sapogenin steroids found in plants into progesterone. His source for the sapogenins was a yam grown in Mexico. Marker and colleagues formed a company (Syntex) to produce progesterone. In 1949, he left the company over financial disputes and destroyed his notes and records. However, a young scientist hired that same year by Syntex ultimately figured prominently in further development of “the Pill.” The new hire, Djerassi, first worked on the synthesis of cortisone from diosgenin. He later turned his attention to synthesizing an “improved” progesterone, one that could be taken orally. In 1951, his group developed a progesterone-like compound called norethindrone. Pincus had been experimenting with the use of progesterone in rabbits to prevent fertility. He ran into an old acquaintance in 1952, John Rock, a gynecologist, who had been using progesterone to enhance fertility in patients who were unable to conceive. Rock theorized that if ovulation were turned off for a short time, the reproductive system would rebound. Rock had essentially proved in humans that progesterone did prevent ovulation. Once Pincus and Rock learned of norethindrone, the stage was set for wider clinical trials that eventually led to FDA approval of it in 1960 as an oral contraceptive. However, many groups opposed this approval on moral, ethical, legal, and religious grounds. Despite such opposition, the Pill was widely used and came to have a profound impact on society. DNA et al. Nucleic acid chemistry and biology were especially fruitful in the 1950s as not only the structure of DNA but also the steps in its replication, transcription, and translation were revealed. In 1952, Rosalind E. Franklin at King’s College in England began producing the X-ray diffraction images that were ultimately used by James Watson and Francis Crick in their elucidation of the structure of DNA published in Science in 1953. Two years later, Severo Ochoa at New York University School of Medicine discovered the enzyme, RNA polymerase, that made RNA from DNA. In 1956, electron microscopy was used to determine that the cellular structures called microsomes contained RNA (they were thus renamed ribosomes). That same year, Arthur Kornberg at Washington University Medical School (St. Louis, MO) discovered DNA polymerase. Soon after that, DNA replication as a semiconservative process was worked out separately by autoradiography in 1957 and then by using density centrifugation in 1958. With the discovery of transfer RNA (tRNA) in 1957 by Mahlon Bush Hoagland at Harvard Medical School, all of the pieces were in place for Francis Crick to postulate in 1958 the “central dogma” of DNA—that genetic information is maintained and transferred in a one-way process, moving from nucleic acids to proteins. The path was set for the elucidation of the genetic code the following decade. On a related note, in 1958, bacterial transduction was discovered by Joshua Lederberg at the University of Wisconsin—a critical step toward future genetic engineering. Probing proteins Behind the hoopla surrounding the discovery of the structure of DNA, the blossoming of protein chemistry in the 1950s is often ignored. Fundamental breakthroughs occurred in the analysis of protein structure and the elucidation of protein functions. In the field of nutrition, in 1950 the protein-building role of the essential amino acids was demonstrated. Linus Pauling, at the California Institute of Technology, proposed that protein structures are based on a primary alpha-helix (a structure that served as inspiration for helical models of DNA). Frederick Sanger at the Medical Research Council (MRC) Unit for Molecular Biology at Cambridge and Pehr Victor Edman developed methods for identifying N-terminal peptide residues, an important breakthrough in improved protein sequencing. In 1952, Sanger used paper chromatography to sequence the amino acids in insulin. In 1953, Max Perutz and John Kendrew, cofounders of the MRC Unit for Molecular Biology, determined the structure of hemoglobin using X-ray diffraction. In 1954, Vincent du Vigneaud at Cornell University synthesized the hormone oxytocin—the first naturally occurring protein made with the exact makeup it has in the body. The same year, ribosomes were identified as the site of protein synthesis. In 1956, the three-dimensional structure of proteins was linked to the sequence of its amino acids, so that by 1957, John Kendrew was able to solve the first three-dimensional structure of a protein (myoglobin); this was followed in 1959 with Max Perutz’s determination of the three-dimensional structure of hemoglobin. Ultimately, linking protein sequences with subsequent structures permitted development of structure–activity models, which allowed scientists to determine the nature of ligand binding sites. These developments proved critical to functional analysis in basic physiological research and to drug discovery, through specific targeting. On to the Sixties By the end of the 1950s, all of pharmaceutical science had been transformed by a concatenation of new instruments and new technologies—from GCs to X-ray diffraction, from computers to tissue culture—coupled, perhaps most importantly, to a new understanding of the way things (meaning cells, meaning bodies) worked. The understanding of DNA’s structure and function—how proteins are designed and how they can cause disease—provided windows of opportunity for drug development that had never before been possible. It was a paradigm shift toward physiology-based medicine, born with the hormone and vitamin work in the 1920s and 1930s, catalyzed by the excitement of the antibiotic era of the 1940s, that continued throughout the rest of the century with the full-blown development of biotechnology-based medicine. The decade that began by randomly searching for antibiotics in dirt ended with all the tools in place to begin searching for drugs with a knowledge of where they should fit in the chemical world of cells, proteins, and DNA. Anodynes & estrogens (1960s) Mention the Sixties and there are varied “hot-button” responses—“JFK, LBJ, civil rights, and Vietnam” or “sex, drugs, and rock ’n’ roll.” But it was all of a piece. Politics and culture mixed like the colors of a badly tie-dyed t-shirt. But in this narrative, drugs are the hallmark of the era—making them, taking them, and dealing with their good and bad consequences. Ever after this era, the world would be continually conscious of pills, pills, pills—for life, for leisure, and for love. In many ways, the Sixties was the Pharmaceutical Decade of the Pharmaceutical Century. A plethora of new drugs was suddenly available: the Pill was first marketed; Valium and Librium debuted to soothe the nerves of housewives and businessmen; blood-pressure drugs and other heart-helping medications were developed. Another emblem of the 1960s was the development of worldwide drug abuse, including the popularization of psychotropic drugs such as LSD by “gurus” like Timothy Leary. The social expansion of drugs for use and abuse in the 1960s forever changed not only the nature of medicine but also the politics of nations. The technology of drug discovery, analysis, and manufacture also proliferated. New forms of chromatography became available, including HPLC, capillary GC, GC/MS, and the rapid expansion of thin-layer chromatography techniques. Proton NMR was developed to analyze complex biomolecules. By the end of the decade, amino acid analyzers were commonplace, and the ultracentrifuge was fully adapted to biomedical uses. Analytical chemistry and biology joined as never before in the search for new drugs and analysis of old ones. And of equal importance, a new progressivism took the stage, especially in the United States, where increasing demand for access to health care and protection from unsafe and fraudulent medications once more led to an increased federal presence in the process of drug development, manufacture, and sale. Popping the Pill If there was any single thing that foreshadowed the tenor of the decade, one medical advance was paramount: In 1960 the first oral contraceptive was mass-marketed. Sex and drugs were suddenly commingled in a single pill, awaiting only the ascendance of rock ’n’ roll to stamp the image of the decade forever. Born of the Pill, the sexual revolution seemed to be in good measure a pharmaceutical one. The earlier achievements of women’s suffrage and the growing presence of women in the labor force were somewhat blocked from further development by the demands of pregnancy and child rearing in a male-dominated society. Feminist historians recount how hopes were suddenly energized by the ability of women to control their own bodies chemically and thus, socially. Chemical equality, at least for those who could afford it and, in the beginning, find the right doctors to prescribe it, was at last available. However, it was not a uniformly smooth process. First and foremost, despite its popularity, the technology for tinkering with women’s reproductive systems had not been fully worked out. In Britain in 1961, there were problems with birth control pills with excess estrogens. Similar problems also occurred in the United States. Dosage changes were required for many women; side effects debilitated a few. But still the sexual revolution marched on, as was documented in the 1966 Masters and Johnson report Human Sexual Response, which showed a transformation of female sexuality, new freedoms, and new attitudes in both sexes. Furthermore, technology was not done fiddling with reproduction by any means. In 1969, the first test-tube fertilization was performed. Valium of the dolls In an era that would be far from sedate, the demand for sedatives was profound, and the drug marketplace responded rapidly. Although Miltown (meprobamate), the first of the major “tranks,” was called the Wonder Drug of 1954, sedatives weren’t widely used until 1961, when Librium (a benzodiazepine) was discovered and marketed as a treatment for tension. Librium proved a phenomenal success. Then Valium (diazepam), discovered in 1960, was marketed by Roche Laboratory in 1963 and rapidly became the most prescribed drug in history. These drugs were touted to the general population and mass-marketed and prescribed by doctors with what many claimed was blithe abandon. While the youth of America were smoking joints and tripping on acid, their parents’ generation of businessmen and housewives were downing an unprecedented number of sedatives. According to the Canadian Government Commission of Inquiry into the Nonmedical Use of Drugs (1972), “In 1965 in the USA, some 58 million new prescriptions and 108 million refills were written for psychotropes (sedatives, tranquilizers, and stimulants), and these 166 million prescriptions accounted for 14% of the total prescriptions of all kinds written in the United States.” Physical and psychological addiction followed for many. Drug taking became the subject of books and movies. In the 1966 runaway best-seller Valley of the Dolls by Jacqueline Susann, the “dolls” were the pills popped by glamorous upper-class women in California. Eventually, the pills trapped them in a world of drug dependence that contributed to ruining their lives. Drug wars It was only a matter of time before the intensive testing of LSD by the military in the 1950s and early 1960s—as part of the CIA’s “Project MKULTRA”—spread into the consciousness of the civilian population. By 1966, the chairman of the New Jersey Narcotic Drug Study Commission declared that LSD was “the greatest threat facing the country today… more dangerous than the Vietnam War.” In the United States, at least at the federal level, the battle against hallucinogens and marijuana use was as intense as, if not more intense than, the fight against narcotic drugs. According to some liberal critics, this was because these “recreational” drugs were a problem in the middle and the upper classes, whereas narcotic addiction was the purview of the poor. From the start of the decade, in the United States and around the world, regions with large populations of urban poor were more concerned about the growing problem of narcotic drug addiction. In 1961, with the passage of the UN Single Convention on Narcotic Drugs, signatory nations agreed to processes for mandatory commitment of drug users to nursing homes. In 1967, New York State established a Narcotics Addiction Control Program that, following the UN Convention, empowered judges to commit addicts into compulsory treatment for up to five years. The program cost $400 million over just three years but was hailed by Governor Nelson Rockefeller as the “start of an unending war.” Such was the measure of the authorities’ concern with seemingly out-of-control drug abuse. The blatant narcotics addictions of many rock stars and other celebrities simultaneously horrified some Americans and glamorized the use of “hard” drugs among others, particularly young people. By 1968, the Food and Drug Administration (FDA) Bureau of Drug Abuse Control and the Treasury Department’s Bureau of Narcotics were fused and transferred to the Department of Justice to form the Bureau of Narcotics and Dangerous Drugs in a direct attempt to consolidate the policing of traffic in illegal drugs. Also in 1968, Britain passed the Dangerous Drug Act to regulate opiates. As these efforts to stop drug use proliferated, technology for mass-producing many of these same drugs continued to improve in factories around the world. By 1968, Canada was producing nearly 56 million doses of amphetamines; by 1969, the United States was producing more than 800,000 pounds of barbiturates. Forging a great society The 1960 election of a Democratic administration in the United States created a new demand for government intervention in broader areas of society. John F. Kennedy and his successor, Lyndon B. Johnson, expanded federal intervention in a host of previously unregulated areas, both civil and economic, including food and medicine. Around the world, a new social agenda demanding rights for minorities (especially apparent in the civil rights struggles in the United States) and women (made possible by freedoms associated with the Pill) fostered a new focus on protecting individuals and ending, or at least ameliorating, some of the damage done to hitherto exploited social classes. Health and medicine, a prime example, led to what many disparagingly called “government paternalism.” This same liberal agenda, in its darker moments, moved to protect less-developed nations from the perceived global communist threat—hence the Vietnam war. Ultimately, there was neither the money nor the will to finance internal social and external political agendas. General inflation resulted, with specific increases in the cost of medical care. Perhaps one of the most significant long-term changes in drug development procedures of the era, especially regarding the role of governments, came with a new desire to protect human “guinea pigs.” General advocacy for the poor, women, and minorities led to a reexamination of the role of the paternalistic, (generally) white male clinicians in the morally repugnant treatment of human subjects throughout the century, before and after the Nuremberg trials of Nazi doctors. It was a profound catalyst for the modern bioethics movement when health groups worldwide established new regulations regarding informed consent and human experimentation in response to specific outrages. In 1962, the Kefauver–Harris amendments to the Federal Food, Drug, and Cosmetic Act of 1938 were passed to expand the FDA’s control over the pharma and food industries. The Kefauver amendments were originally the outgrowth of Senate hearings begun in 1959 to examine the conduct of pharmaceutical companies. According to testimony during those hearings, it was common practice for these companies “to provide experimental drugs whose safety and efficacy had not been established to physicians who were then paid to collect data on their patients taking these drugs. Physicians throughout the country prescribed these drugs to patients without their control or consent as part of this loosely controlled research.” That the amendments were not passed until 1962 was in part because of the profound battle against allowing additional government control of the industry. The 1958 Delaney proviso and the 1960 Color Additive Amendment led to industry and conservative resentment and complaints that the FDA was gaining too much power. However, with the 1961 birth defects tragedy involving the popular European sedative thalidomide (prevented from being marketed in the United States by FDA researcher Frances Kelsey), public demand for greater protections against experimental agents was overwhelming. Thalidomide had been prescribed to treat morning sickness in countless pregnant women in Europe and Canada since 1957, but its connection to missing and deformed limbs in newborns whose mothers had used it was not realized until the early 1960s. In 1961, televised images of deformed “thalidomide babies” galvanized support for the FDA, demonstrating the power of this new medium to transform public opinion, as it had in the 1960 Nixon–Kennedy debates and would again in the Vietnam War years. Building on the newfound public fears of synthetic substances, Rachel Carson’s 1962 book Silent Spring precipitated the populist environmental movement. As with the 1958 Delaney clause and the Federal Hazardous Substances Labeling Act of 1960, it was all part of increased public awareness of the potential dangers of the rapid proliferation of industrial chemicals and drugs. According to the 1962 amendments to the 1938 Food, Drug, and Cosmetic Act, new drugs had to be shown to be effective, prior to marketing, by means to be determined by the FDA. Ultimately, this requirement translated to defined clinical trials. However, Congress still vested excessive faith in the ethics of physicians by eliminating the need for consent when it was “not feasible” or was deemed not in the best interest of the patient; these decisions could be made “according to the best judgment of the doctors involved.” Despite the fact that President Kennedy proclaimed a Consumer Bill of Rights—including the rights to safety and to be informed—more stringent federal guidelines to protect research subjects were not instituted until 1963. Then the NIH responded to two egregious cases: At Tulane University, a chimpanzee kidney was unsuccessfully transplanted into a human with the patient’s consent but without medical review. At the Brooklyn Jewish Chronic Disease Hospital, in association with the Sloan-Kettering Cancer Research Institute, live cancer cells were injected into indigent, cancer-free elderly patients. Particularly important to the public debate was the disclosure of 22 examples of potentially serious ethical violations found in research published in recent medical journals. The information was presented by Henry Beecher of Harvard Medical School to science journalists at a 1965 conference sponsored by the Upjohn pharmaceutical company and was later published in the New England Journal of Medicine. In 1964, the World Medical Association issued its Declaration of Helsinki, which set standards for clinical research and demanded that subjects be given informed consent before enrolling in an experiment. By 1966, in the United States, the requirement for informed consent and peer review of proposed research was written into the guidelines for Public Health Service–sponsored research on human subjects. These regulations and the debate surrounding human experimentation continued to evolve and be applied more broadly. And in an attempt to protect consumers from unintentional waste or fraud, in 1966, wielding its new power from the Kefauver amendments, the FDA contracted with the National Academy of Sciences/National Research Council to evaluate the effectiveness of 4000 drugs approved on the basis of safety alone between 1938 and 1962. The U.S. government faced other health issues with less enthusiasm. In 1964, the Surgeon General’s Report on Smoking was issued, inspired in part by the demand for a response to the Royal College of Surgeons’ report in Britain from the year before. Although the Surgeon General’s report led to increased public awareness and mandated warning labels on tobacco products, significant government tobacco subsidies, tax revenues, smoking, and smoking-related deaths continued in the United States. The debate and associated litigation go on to this day. Heart heroics and drugs The lungs were not the only vital organs to attract research and publicity in the Sixties. From heart transplants and blood-pressure medications to blood-based drugs, a new kind of “heroic” medicine focused technology on the bloodstream and its potential for helping or hindering health. Heart surgery captivated the public with its bravado. In 1960, the transistorized self-contained pacemaker was introduced. In the 1960s, organ transplantation became routinely possible with the development of the first immune suppressant drugs. In 1967, the coronary bypass operation was developed by Michael DeBakey; in that same year, physician (and some would say showman) Christiaan Barnard performed the first heart transplant. In 1968, angioplasty was introduced for arterial treatment and diagnosis. But for all the glamour of surgery, chemicals for controlling heart disease provided far more widespread effects. Blood-pressure drugs based on new knowledge of human hormone systems became available for the first time. In 1960, guanethidine (a noradrenaline release inhibitor) was developed for high blood pressure; in rapid succession the first beta-adrenergic blocker appeared in Britain (1962), and alpha-methyldopa, discovered in 1954, was first used clinically in the early 1960s for treating high blood pressure by interfering not with noradrenaline release, but with its synthesis. In 1964, methods were perfected to nourish individuals through the bloodstream. This ability to feed intravenously and provide total caloric intake for debilitated patients forever changed the nature of coma and the ethics of dealing with the dying. The use of blood-thinning and clot-dissolving compounds for heart disease was also pioneered in the early 1960s: Streptokinase and aspirin reduced deaths by 40% when taken within a few hours of a heart attack. In the late Sixties, as methods of fractionation using centrifugation and filtration improved dramatically, concentrated blood factors became readily available for the first time. Plasmapheresis—centrifuging the red blood cells from plasma and then returning the whole cells to the donor—allowed donations to occur twice weekly instead of every few months. Blood proteins such as “Big-D” Rh antibodies used to immunize women immediately after the birth of their first Rh-positive child, albumin, gamma globulins, blood-typing sera, and various clotting factors such as factor VIII created a booming industry in plasma products. But it also made blood all the more valuable as a commodity. Abuses increased in collecting, distributing, and manufacturing blood products, as recounted by Douglas Starr in his 1999 book Blood. In 1968, in a move that foreshadowed the AIDS era yet to come, the United States revoked all licenses for consumer sales of whole plasma prepared from multiple donors because of fears that viral hepatitis would spread. Fears were exacerbated by scandalous revelations about the health status (or lack thereof) of many paid blood donors. This donor problem became even more newsworthy in the early 1970s as malnourished street hippies, drug abusers, unhealthy indigents, and prisoners were revealed as sources often contaminating the world blood supply. High tech/new mech In the 1960s, analytical chemists increasingly turned their attention to drug discovery and analysis and to fundamental questions of biological significance in physiology and genetics. New technologies were developed, and previously designed instruments were adapted to biomedical applications. The development of high-pressure (later known as high-performance) liquid chromatography (HPLC) heralded a new era of biotechnology and allowed advanced separations of fragile macromolecules in fractions of the time previously required. The radioimmune assay, first developed in 1959 by Rosalyn Yalow and Solomon Berson, was perfected in 1960. Tissue culture advances proliferated, allowing more and better in vitro testing of drugs; and, when coupled with radiotracer and radioimmunoassay experiments, led to unprecedented breakthroughs in all areas of mammalian physiology. In 1964, for example, Keith Porter and Thomas F. Roch discovered the first cell membrane receptor. Further developments in analytical chemistry came as gas chromatography (GC) was first linked with mass spectrometry (MS), providing a quantum leap in the ability to perform structural analysis of molecules. And laboratory automation, including primitive robotics, became a powerful trend. But perhaps the most important breakthrough in tissue culture, and one that created a direct path to the Human Genome Project, was the invention of somatic-cell hybridization by Mary Weiss and Howard Green in 1965. By fusing mouse and human cells together via the molecular “glue” of Sendai virus, these researchers and others quickly developed a series of cell lines containing mostly mouse chromosomes but with different single human ones, all expressing unique proteins. For the first time, human proteins could be assigned to individual human chromosomes (and later chromosome fragments) to the degree that gene mapping was finally possible in humans. Of comparable importance, new improvements in fermentation technology allowed continuous cycling and easy sterilization along with mass-produced instrumentation. Fundamental breakthroughs in the etiology of disease transformed the understanding of infection. In 1961, the varying polio virus receptors were correlated to pathogenicity of known isolates; in 1967, diphtheria toxin’s mode of action was finally determined and provided the first molecular definition of a bacterial protein virulence factor. Structural biology proceeded at an enthusiastic pace. In 1960, John Kendrew reported the first high-resolution X-ray analysis of the three-dimensional structure of a protein—sperm whale myoglobin. In the 1960s, image analyzers were linked to television screens for the first time, enhancing the use and interpretation of complex images. And in 1967, Max Perutz and Hilary Muirhead built the first high-resolution model of the atomic structure of a protein (oxyhemoglobin), which promoted a wave of protein structural analysis. Computer systems became increasingly powerful and quickly indispensable to all laboratory processes, but especially to molecular analysis. Hardly falling under the category of miscellaneous discoveries, at least in terms of the development of biotechnology and the pharmaceutical industry, was the development of agarose gel electrophoresis in 1961. This requirement was critical for separating and purifying high molecular weight compounds, especially DNA; in 1963, to the benefit of a host of laboratory workers, the first film badge dosimeter was introduced in the United Kingdom; and in 1966, the disc-diffusion standardized test was developed for evaluating antibiotics, which was a boon to the exploding pharmaceutical development of such compounds. Dancing with DNA Finally, the understanding of the structure of DNA and acceptance that it was indeed genetic material yielded intelligible and practical results. At the beginning of the decade, researchers A. Tsugita and Heinz Fraenkel-Conrat demonstrated the link between mutation and a change in the protein produced by the gene. Also in 1960, Francois Jacob and Jacques Monod proposed their operon model. This was the birth of gene regulation models, which launched a continuing quest for gene promoters and triggering agents, such that by 1967, Walter Gilbert and co-workers identified the first gene control (repressor) substance. Perhaps most significantly, the prelude to biotechnology was established when restriction enzymes were discovered, the first cell-free DNA synthesis was accomplished by biochemist Arthur Kornberg in 1961, and the DNA–amino acid code was deciphered. Deciphering the genetic code was no minor feat. Finding the mechanism for going from gene to protein was a combination of brilliant theorizing and technological expertise. The technology necessary for the dawn of biotechnology proliferated as well. Throughout the 1960s, automated systems for peptide and nucleic acid analysis became commonplace. In 1964, Bruce Merrifield invented a simplified technique for protein and nucleic acid synthesis, which was the basis for the first such machines. (In the 1980s, this technique would be mass-automated for gene synthesis.) And in 1967, the first specific gene transfer was accomplished; the lac operon was functionally transferred from E. coli to another bacterial species. And, although its importance was not realized at the time, in 1967, Thermus aquaticus was discovered in a hot spring in Yellowstone National Park. This microbe was the first Archea ever found and the ultimate source of heat-stable taq polymerase—the enabling enzyme for modern PCR. By the late 1960s, not only had the first complete gene been isolated from an organism, but biologists were already debating the ethics of human and animal genetic engineering. Generation of drug-takers In the 1960s, the post–World War II generation of baby boomers entered their teenage years. These were the children of vitamins, antibiotics, hormones, vaccines, and fortified foods such as Wonder Bread and Tang. The technology that won the war had also transformed the peace and raised high expectations on all social fronts, including, and perhaps especially, health. The unparalleled prosperity of the 1950s was largely driven by new technologies and a profusion of consumer goods. This prosperity, coupled with a new political progressivism, created a euphoric sense of possibility early in the decade, especially with regard to health care. The prevailing belief was that medicine would save and society would provide. Although the dream ultimately proved elusive, its promise permeated the following decades—Medicare, Medicaid, and a host of new regulations were the lingering aftereffects in government, a generation of drug-takers the result in society at large. The Sabin oral polio vaccine was approved in the United States in 1960 after trials involving 100 million people overseas and promised salvation to a new generation from this former scourge. In 1961, the sweetener cyclamate was introduced in the first low-calorie soft drink, Diet Rite, and it created demand for consumption without cost, or at least without weight gain. In 1964, a suitable, routine vaccine for measles was developed that was much better than its predecessor vaccine first produced in 1960. In 1967, a live-virus mumps vaccine was developed. Faith that childhood diseases could be stamped out grew rapidly. The Surgeon General of the United States even went so far as to state that we were coming near—for the first time in history—to finally “closing the books on infectious diseases.” From today’s perspective, these seem like naive hopes. But they were not without some foundation. Antibiotics had yet to lose effectiveness, and more were being discovered or synthesized. In 1966, the first antiviral drug, amantadine-HCl, was licensed in the United States for use against influenza. In 1967, the who began efforts to eradicate smallpox. The rhetoric of nongovernmental antipolio and antituberculosis groups provided additional reason for optimism. No wonder a generation of baby boomers was poised to demand a pill or vaccine for any and every ill that affected humankind—pharmaceutical protection from unwanted pregnancies, from mental upset, from disease. Rising expectations were reflected in the science and business arenas, where research was promoted and industry driven to attack a greater variety of human ills with the weapons of science. Technological fixes in the form of pills and chemicals seemed inevitable. How else could the idea of a war on cancer in the next decade be initiated with the same earnestness and optimism as the quest for a man on the moon? The 1969 success of Apollo 11 was the paradigm for the capabilities of technology. On a darker note, the decade ended with a glimpse of the more frightening aspects of what biomedical technology could do, or at least what militarists wanted it to do. In 1969, the U.S. Department of Defense requested $10 million from Congress to develop a synthetic biological agent to which no natural immunity existed. Funding soon followed under the supervision of the CIA at Fort Detrick, MD. Ultimately, although a new battery of technologies had become available in the 1960s, including HPLC, GC/MS, and machines to synthesize DNA and proteins, and new knowledge bases were developed—from cracking the genetic code to the discovery of restriction enzymes—the majority of these breakthroughs would not bear fruit until the 1970s and 1980s. The 1960s would instead be remembered primarily for the changes wrought by new pharmaceuticals and new social paradigms. As the decade ended, it was an open question as to whether the coming decade would bring the dawn of massive biological warfare or the rather expected nuclear holocaust (heightened in the world psyche by the Cuban Missile Crisis of 1962). Others speculated that the future would bring massive social collapse arising from the intergenerational breakdown in the West that many blamed on the success of pharmaceutical technology in developing and producing new and dangerous drugs. Chemistry, cancer & ecology (1970s) As the 1970s opened, new chemistries and the war on cancer seized center stage. U.S. President Richard Nixon (taking a moment of from his pursuit of the Vietnam War) established the National Cancer Program, popularly known as the war on cancer, with an initial half-billion dollars of new funding. Carcinogens were one of the concerns in the controversy surrounding the polluted Love Canal. And cancer was especially prominent in the emotional issue of the “DES daughters”—women at risk for cancer solely because of diethylstilbestrol (DES), the medication prescribed to their mothers during pregnancy. New cancer treatments were developed; chemotherapy joined the ranks of routine treatments, especially for breast cancer. New drugs appeared. Cyclosporin provided a long-sought breakthrough with its ability to prevent immune rejection of tissue grafts and organ transplants. Rifampicin proved its worth for treating tuberculosis; cimetidine (Tagamet), the first histamine blocker, became available for treating peptic ulcers. Throughout the decade, improvements in analytical instrumentation, including high-pressure liquid chromatography (HPLC) and mass spectrometry, made drug purification and analysis easier than ever before. In this period, NMR became transformed into the medical imaging system, MRI. The popular environmental movement that took root in the ideology of the previous decade blossomed politically in 1970 as the first Earth Day was celebrated, the U.S. Clean Air Act was passed, and the U.S. Environmental Protection Agency (EPA) was established. Some of the optimism of the Sixties faded as emerging plagues, such as Lyme and Legionnaires’ disease in the United States and Ebola and Lassa fever in Africa, reopened the book on infectious diseases. The World Health Organization continued its smallpox eradication campaign, but as DDT was gradually withdrawn because of its detrimental effect on the environment, efforts to eradicate malaria and sleeping sickness were imperiled. Ultimately, the 1970s saw the start of another kind of infection—genetic engineering fever—as recombinant DNA chemistry dawned. In 1976, in a move to capitalize on the new discoveries, Genentech Inc. (San Francisco) was founded and became the prototypical entrepreneurial biotech company. The company’s very existence forever transformed the nature of technology investments and the pharmaceutical industry. Cancer wars The first major salvo against cancer in the United States was in 1937, when the National Cancer Institute (NCI) was created by congressional mandate. A sign of the times, one of the key provisions of the act was to enable the institute to “procure, use, and lend radium.” Such an early interest in cancer was by no means unique to the United States. Throughout the period, Nazi Germany led the world in cancer research, including early demonstrations of the carcinogenic effects of smoking tobacco. Cancer remained of interest to researchers throughout the world in the decades to follow. In the 1950s, a major move was made to develop chemotherapies for various cancers. By 1965, the NCI had instituted a program specifically for drug development with participation from the NIH, industry, and universities. The program screened 15,000 new chemicals and natural products each year for potential effectiveness. Still, by the 1970s, there seemed to be a harsh contrast between medical success against infectious diseases and cancer. The 1971 report of the National Panel of Consultants on the Conquest of Cancer (called the Yarborough Commission) formed the basis of the 1971 National Cancer Act signed by President Nixon. The aim of the act was to make “the conquest of cancer a national crusade” with an initial financial boost of $500 million (which was allocated under the direction of the long-standing NCI). The Biomedical Research and Research Training Amendment of 1978 added basic research and prevention to the mandate for the continuing program. Daughters and sons The story of the synthetic estrogen, DES, framed cancer as a complex societal problem and not “just” an issue dealing with a particular new source of cancer. DES daughters not only crossed generations, but also involved interactions between patients, the drug industry, uninformed or wrongly informed physicians, and the political interests of Congress and the FDA. DES was prescribed from the early 1940s until 1971 to help prevent certain complications of pregnancy, especially those that led to miscarriages. By the 1960s, DES use was decreasing because of evidence that the drug lacked effectiveness and might indeed have damaging side effects, although no ban or general warning to physicians was issued. According to the University of Pennsylvania Cancer Center (Philadelphia), there are few reliable estimates of the number of women who took DES, although one source estimates that 5–10 million women either took the drug during pregnancy or were exposed to it in utero. In 1970, a study in the journal Cancer described a rare form of vaginal cancer, clear cell adenocarcinoma (CCAC). The following year, a study in The New England Journal of Medicine documented the association between in utero DES exposure and the development of CCAC. By the end of that year, the FDA issued a drug bulletin warning of potential problems with DES and advised against its use during pregnancy. So-called DES daughters experienced a wide variety of effects including infertility, reproductive tract abnormalities, and increased risks of vaginal cancer. More recently, a number of DES sons were also found to have increased levels of reproductive tract abnormalities. In 1977, inspired by the tragedies caused by giving thalidomide and DES to pregnant women, the FDA recommended against including women of child-bearing potential in the early phases of drug testing except for life-threatening illnesses. The discovery of DES in beef from hormone-treated cattle after the FDA drug warning led in 1979 to what many complained was a long-delayed ban on its use by farmers. The DES issue was one of several that helped focus part of the war against cancer as a fight against environmental carcinogens (see below). Cancer research/cancer “cures” New evidence at the beginning of the decade seemed to promote an infectious model of cancer development. In 1970, Howard Martin Temin (at the University of Wisconsin– Madison) and David Baltimore (at the Massachusetts Institute of Technology; MIT) independently discovered viral reverse transcriptase, showing that some RNA viruses (the retroviruses), in their own version of genetic engineering, were capable of creating DNA copies of themselves. The viral DNA was able to integrate into the infected host cell, which then transformed into a cancer cell. Reverse transcriptase eventually became critical to the study of the aids virus in following decades. Temin and Baltimore shared the 1975 Nobel Prize in Physiology or Medicine with Renato Dulbecco of the Salk Institute. Many claims of cancer virus discoveries based on animal studies came and went early in the decade, but they proved inapplicable to humans. Hopes of treating the class of diseases known as cancers with traditional vaccination techniques declined. Other research developments helped expand knowledge of the mechanics and causes of cancer. In 1978, for example, the cancer suppressor gene P53 was first observed by David Lane at the University of Dundee. By 1979, it was possible to use DNA from malignant cells to transform cultured mouse cells into tumors—creating an entirely new tool for cancer study. Although many treatments for cancer existed at the beginning of the 1970s, few real cures were available. Surgical intervention was the treatment of choice for apparently defined tumors in readily accessible locations. In other cases, surgery was combined with or replaced by chemotherapy and/or radiation therapy. Oncology remained, however, as much an art as a science in terms of actual cures. Too much depended on too many variables for treatments to be uniformly applicable or uniformly beneficial. There were, however, some obvious successes in the 1970s. Donald Pinkel of St. Jude’s Hospital (Memphis) developed the first cure for acute lymphoblastic leukemia, a childhood cancer, by combining chemotherapy with radiotherapy. The advent of allogenic (foreign donor) bone marrow transplants in 1968 made such treatments possible, but the real breakthrough in using powerful radiation and chemotherapy came with the development of autologous marrow transplantation. The method was first used in 1977 to cure patients with lymphoma. Autologous transplantation involves removing and usually cryopreserving a patient’s own marrow and reinfusing that marrow after the administration of high-dosage drug or radiation therapy. Because autologous marrow can contain contaminating tumor cells, a variety of methods have been established to attempt to remove or deactivate them, including antibodies, toxins, and even in vitro chemotherapy. E. Donnall Thomas of the Fred Hutchinson Cancer Research Center (Seattle) was instrumental in developing bone marrow transplants and received the 1990 Nobel Prize in Physiology or Medicine for his work. Although bone marrow transplants were originally used primarily to treat leukemias, by the end of the century, they were used successfully as part of high-dose chemotherapy regimes for Hodgkin’s disease, multiple myeloma, neuroblastoma, testicular cancer, and some breast cancers. In 1975, a WHO survey showed that death rates from breast cancer had not declined since 1900. Radical mastectomy was ineffective in many cases because of late diagnosis and the prevalence of undetected metastases. The search for alternative and supplemental treatments became a high research priority. In 1975, a large cooperative American study demonstrated the benefits of using phenylalanine mustard following surgical removal of the cancerous breast. Combination therapies rapidly proved even more effective; and by 1976, CMF (cyclophosphamide, methotrexate, and 5-fluorouracil) therapy was developed at the Instituto Nazionale Tumori in Milan, Italy. It proved to be a radical improvement over surgery alone and rapidly became the chemotherapy of choice for this disease. A new environment Launched in part by Rachel Carson’s book, Silent Spring, in the previous decade, the environmental movement in the West became ever more prominent. The first Earth Day was held on April 22, 1970, to raise environmental awareness. The EPA was launched on December 2, and Nixon closed out the year by signing the Clean Air Act on December 31. The concept of carcinogens entered the popular consciousness. Ultimately, the combination of government regulations and public fears of toxic pollutants in food, air, and water inspired improved technologies for monitoring extremely small amounts of chemical contaminants. Advances were made in gas chromatography, ion chromatography, and especially the EPA-approved combination of GC/MS. The 1972 Clean Water Act and the Federal Insecticide and Rodenticide Act added impetus to the need for instrumentation and analysis standards. By the 1980s, many of these improvements in analytical instrumentation had profound effects on the scientific capabilities of the pharmaceutical industry. One example is atomic absorption spectroscopy, which in the 1970s made it possible to assay trace metals in foods to the parts-per-billion range. The new power of such technologies enabled nutritional researchers to determine, for the first time, that several trace elements (most usually considered pollutants) were actually necessary to human health. These included tin (1970), vanadium (1971), and nickel (1973). In 1974, the issue of chemical-induced cancer became even broader when F. Sherwood Rowland of the University of California–Irvine and Mario Molina of MIT demonstrated that chlorofluorocarbons (CFCs) such as Freon could erode the UV-absorbing ozone layer. The predicted results were increased skin cancer and cataracts, along with a host of adverse effects on the environment. This research led to a ban of CFCs in aerosol spray cans in the United States. Rowland and Molina shared the 1995 Nobel Prize in Chemistry for their ozone work with Paul Crutzen of the Max Planck Institute for Chemistry (Mainz, Germany). By 1977, asbestos toxicity had become a critical issue. Researchers at the Mount Sinai School of Medicine (New York) discovered that asbestos inhalation could cause cancer after a latency period of 20 years or more. This discovery helped lead to the passage of the Toxic Substances Control Act, which mandated that the EPA inventory the safety of all chemicals marketed in the United States before July 1977 and required manufacturers to provide safety data 90 days before marketing any chemicals produced after that date. Animal testing increased where questions existed, and the issue of chemical carcinogenicity became prominent in the public mind and in the commercial sector. Also in the 1970s, DDT was gradually withdrawn from vector eradication programs around the world because of the growing environmental movement that resulted in an outright ban on the product in the United States in 1971. This created a continuing controversy, especially with regard to who attempts to eliminate malaria and sleeping sickness in the developing world. In many areas, however, the emergence of DDT-resistant insects already pointed to the eventual futility of such efforts. Although DDT was never banned completely except by industrialized nations, its use declined dramatically for these reasons. Recombinant DNA and more In 1970, two years before the birth of recombinant DNA (rDNA) technology, cytogeneticist Robert John Cecil Harris coined the term “genetic engineering.” But more importantly, in 1970, Werner Arber of the Biozentrum der Universität Basel (Switzerland) discovered restriction enzymes. Hamilton O. Smith at Johns Hopkins University (Baltimore) verified Arber’s hypothesis with a purified bacterial restriction enzyme and showed that this enzyme cuts DNA in the middle of a specific symmetrical sequence. Daniel Nathans, also at Johns Hopkins, demonstrated the use of restriction enzymes in the construction of genetic maps. He also developed and applied new methods of using restriction enzymes to solve various problems in genetics. The three scientists shared the 1978 Nobel Prize in Physiology or Medicine for their work in producing the first genetic map (of the SV40 virus). In 1972, rDNA was born when Paul Berg of Stanford University demonstrated the ability to splice together blunt-end fragments of widely disparate sources of DNA. That same year, Stanley Cohen and Herbert Boyer, both from Stanford, met at a Waikiki Beach delicatessen where they discussed ways to combine plasmid isolation with DNA splicing. They had the idea to combine the use of the restriction enzyme EcoR1 (which Boyer had discovered in 1970 and found capable of creating “sticky ends”) with DNA ligase (discovered in the late 1960s) to form engineered plasmids capable of producing foreign proteins in bacteria—the basis for the modern biotechnology industry. By 1973, Cohen and Boyer had produced their first recombinant plasmids. They received a patent on this technology for Stanford University that would become one of the biggest money-makers in pharmaceutical history. The year 1975 was the year of DNA sequencing. Walter Gilbert and Allan Maxam of Harvard University and Fred Sanger of Cambridge University simultaneously developed different methods for determining the sequence of bases in DNA with relative ease and efficiency. For this accomplishment, Gilbert and Sanger shared the 1980 Nobel Prize in Physiology or Medicine. By 1976, Silicon Valley venture capitalist Robert Swanson teamed up with Herbert Boyer to form Genentech Inc. (short for genetic engineering technology). It was the harbinger of a wild proliferation of biotechnology companies over the next decades. Genentech’s goal of cloning human insulin in Escherichia coli was achieved in 1978, and the technology was licensed to Eli Lilly. In 1977, the first mammalian gene (the rat insulin gene) was cloned into a bacterial plasmid by Axel Ullrich of the Max Planck Institute. In 1978, somatostatin was produced using rDNA techniques. The recombinant DNA era grew from these beginnings and had a major impact on pharmaceutical production and research in the 1980s and 1990s. High tech/new mech In June 1970, Raymond V. Damadian at the State University of New York discovered that cancerous tissue in rats exhibited dramatically prolonged NMR relaxation times. He also found that the relaxation times of normal tissues also vary significantly, although less dramatically than cancer tissue. Damadian’s March 1971 Science article, “Tumor Detection by Nuclear Magnetic Resonance,” became the basis for magnetic resonance imaging (MRI)’s pioneer patent, issued to him in 1974, which included a description of a three-dimensional in vivo method for obtaining diagnostic NMR signals from humans. In 1977, Damadian and colleagues achieved the first NMR image of a human in a whole-body MRI scanner. In 1988, Damadian was awarded the National Medal of Technology. MRI became one of the most sensitive and useful tools for disease diagnosis, and the basis of MR spectroscopy, itself one of the most sophisticated in vivo physiological research tools available by the end of the century. In 1971, the first coaxial tomography scanner was installed in England. By 1972, the first whole-body computed tomography (CT) scanner was marketed by Pfizer. That same year, the Brookhaven Linac Isotope Producer went on line, helping to increase the availability of isotopes for medical purposes and basic research. In 1977, the first use of positron emission tomography (PET) for obtaining brain images was demonstrated. The 1970s saw a revolution in computing with regard to speed, size, and availability. In 1970, Ted Hoff at Intel invented the first microprocessor. In 1975, the first personal computer, the Altair, was put on the market by American inventor Ed Roberts. Also in 1975, William Henry Gates III and Paul Gardner Allen founded Microsoft. And in 1976, the prototype for the first Apple Computer (marketed as the Apple II in 1977) was developed by Stephen Wozniak and Steven Jobs. It signaled the movement of personal computing from the hobbyist to the general public and, more importantly, into pharmaceutical laboratories where scientists used PCs to augment their research instruments. In 1975, Edwin Mellor Southern of the University of Oxford invented a blotting technique for analyzing restriction enzyme digest fragments of DNA separated by electrophoresis. This technique became one of the most powerful technologies for DNA analysis and manipulation and was the conceptual template for the development of northern blotting (for RNA analysis) and western blotting (for proteins). Although HPLC systems had been commercially available from ISCO since 1963, they were not widely used until the 1970s when—under license to Waters Associates and Varian Associates—the demands of biotechnology and clinical practice made such systems seem a vibrant new technology. By 1979, Hewlett-Packard was offering the first microprocessor-controlled HPLC, a technology that represented the move to computerized systems throughout the life sciences and instrumentation in general. Throughout the decade, GC and MS became routine parts of life science research, and the first linkage of LC/MS was offered by Finnigan. These instruments would have increasing impact throughout the rest of the century. (Re)emerging diseases In 1969, U.S. Surgeon General William Stewart claimed in testifying before Congress that “the time has come to close the book on infectious diseases.” He believed that it was especially time to reinvest money to treat killers such as heart disease and cancer, since, in his opinion, it was only a matter of time before the war against infection would be won by a combination of antibiotics and vaccines. The traditional diseases were indeed on the run before the onslaught of modern medicines. What he could not foresee was the emergence of new diseases and the reemergence of old plagues in the guise of drug-resistant strains. The 1970s helped throw cold water on this kind of optimism, perhaps signaled by the shift in emphasis implied by the 1970 name change of the Communicable Disease Center to the Centers for Disease Control. There were certainly enough new diseases and “old friends” to control. For example, in 1972, the first cases of recurrent polyarthritis (Lyme disease) were recorded in Old Lyme and Lyme, CT, ultimately resulting in the spread of the tick-borne disease throughout the hemisphere. The rodent-borne arena viruses were identified in the 1960s and shown to be the causes of numerous diseases seen since 1934 in both developed and less-developed countries. Particularly deadly was the newly discovered Lassa fever virus, first identified in Africa in 1969 and responsible for small outbreaks in 1970 and 1972 in Nigeria, Liberia, and Sierra Leone, with a mortality rate of some 36–38%. Then, in 1976, epidemics of a different hemorrhagic fever occurred simultaneously in Zaire and Sudan. Fatalities reached 88% in Zaire (now known as the Democratic Republic of the Congo) and 53% in Sudan, resulting in a total of 430 deaths. Ebola virus, named after a small river in northwest Zaire, was isolated from both epidemics. In 1977, a fatality was attributed to Ebola in a different area of Zaire. The investigation of this death led to the discovery that there were probably two previous fatal cases. A missionary physician contracted the disease in 1972 while conducting an autopsy on a patient thought to have died of yellow fever. In 1979, the original outbreak site in Sudan generated a new episode of Ebola hemorrhagic fever that resulted in 34 cases with 22 fatalities. Investigators were unable to discover the source of the initial infections. The dreaded nature of the disease—the copious bleeding, the pain, and the lack of a cure—sent ripples of concern throughout the world’s medical community. In 1976, the unknown “Legionnaires’ disease” appeared at a convention in Philadelphia, killing 29 American Legion convention attendees. The cause was identified as the newly discovered bacterium Legionella. Also in 1976, a Nobel Prize in Physiology or Medicine was awarded to Baruch Blumberg (originally at the NIH, then at the University of Pennsylvania) for the discovery of a new disease agent in 1963—hepatitis B, for which he helped to develop a blood test in 1971. Nonetheless, one bit of excellent news at the end of the decade made such “minor” outbreaks of new diseases seem trivial in the scope of human history. From 1350 B.C., when the first recorded smallpox epidemic occurred during the Egyptian–Hittite war, to A.D. 180, when a large-scale epidemic killed between 3.5 and 7 million people (coinciding with the first stages of the decline of the Roman Empire), through the millions of Native Americans killed in the 16th century, smallpox was a quintessential scourge. But in 1979, a WHO global commission was able to certify the worldwide eradication of smallpox, achieved by a combination of quarantines and vaccination. The last known natural case of the disease occurred in 1977 in Somalia. Government stocks of the virus remain a biological warfare threat, but the achievement may still, perhaps, be considered the most unique event in the Pharmaceutical Century—the disease-control equivalent of landing a man on the moon. By 1982, vaccine production ceased. By the 1990s, a controversy arose between those who wanted to maintain stocks in laboratories for medical and genetic research purposes (and possibly as protection against clandestine biowarfare) and those who hoped to destroy the virus forever. On a lesser but still important note, the first leprosy vaccine using the nine-banded armadillo as a source was developed in 1979 by British physician Richard Rees at the National Institute for Medical Research (Mill Hill, London). Toward a healthier world The elimination of smallpox was just part of a large-scale move in the 1970s to deal with the issue of global health. In 1974, the WHO launched an ambitious Expanded Program on Immunization to protect children from polio myelitis, measles, diphtheria, whooping cough, tetanus, and tuberculosis. Throughout the 1970s and the rest of the century, the role of DDT in vector control continued to be a controversial issue, especially for the eradication or control of malaria and sleeping sickness. The WHO would prove to be an inconsistent ally of environmental groups that urged a ban of the pesticide. The organization’s recommendations deemphasized the use of the compound at the same time that its reports emphasized its profound effectiveness. Of direct interest to the world pharmaceutical industry, in 1977 the WHO published the first Model List of Essential Drugs—“208 individual drugs which could together provide safe, effective treatment for the majority of communicable and noncommunicable diseases.” This formed the basis of a global movement toward improved health provision as individual nations adapted and adopted this list of drugs as part of a program for obtaining these universal pharmacological desiderata. That ’70s showdown The decade lurched from the OPEC oil embargoes to Watergate, through stagflation, world famine, and hostage crises to a final realization, in 1979, that it all might have started with a comet or asteroid that crashed into the earth 65 million years before, killing off the dinosaurs. And that perhaps it might end the same way. Medical marvels and new technologies promised excitement at the same time that they revealed more problems to be solved. A new wave of computing arose in Silicon Valley. Oak Ridge National Laboratory detected a single atom for the first time—one atom of cesium in the presence of 1019 argon atoms and 1018 methane molecules—using lasers. And biotechnology was born as both a research program and a big business. The environmental movement contributed not only a new awareness of the dangers of carcinogens, but a demand for more and better analytical instruments capable of extending the range of chemical monitoring. These would make their way into the biomedical field with the demands of the new biotechnologies. The pharmaceutical industry would enter the 1980s with one eye on its pocketbook, to be sure, feeling harried by economics and a host of new regulations, both safety and environmental. But the other eye looked to a world of unimagined possibilities transformed by the new DNA chemistries and by new technologies for analysis and computation. Arteries, AIDS, and Engineering (1980s) Whether the changes to the pharmaceutical industry and the world in the 1980s will prove most notable for the rise of and reaction to a new disease, AIDS, or the flowering of entrepreneurial biotechnology and genetic engineering, it is too soon to say. These changes—along with advances in immunology, automation, and computers, the development of new paradigms of cardiovascular and other diseases, and restructured social mores—all competed for attention in the transformational 1980s. AIDS: A new plague It struck the big cities first, and within those cities, at first, it only affected certain segments of the population, primarily homosexual men. The first published reports of the new disease seemed like no more than medical curiosities. On June 5, 1981, the Atlanta-based Centers for Disease Control and Prevention (CDC), a federal agency charged with keeping tabs on disease, published an unusual notice in its Morbidity and Mortality Weekly Report: the occurrence of Pneumocystis carinii pneumonia (PCP) among gay men. In New York, a dermatologist encountered cases of a rare cancer, Kaposi’s sarcoma (KS), a disease so obscure he recognized it only fromdescriptions in antiquated textbooks. By the end of 1981, PCF and KS were recognized as harbingers of a new and deadly disease. The disease was initially called Gay Related Immune Deficiency. Within a year, similar symptoms appeared in other demographic groups, primarily hemophiliacs and users of intravenous drugs. The CDC renamed the disease Acquired Immune Deficiency Syndrome (AIDS). By the end of 1983, the CDC had recorded some 3000 cases of this new plague. The prospects for AIDS patients were not good: almost half had already died. AIDS did not follow normal patterns of disease and infection. It produced no visible symptoms—at least not until the advanced stages of infection. Instead of triggering an immune response, it insidiously destroyed the body’s own mechanisms for fighting off infection. People stricken with the syndrome died of a host of opportunistic infections such as rare viruses, fungal infections, and cancers. When the disease began to appear among heterosexuals, panic and fear increased. Physicians and scientists eventually mitigated some of the hysteria when they were able to explain the methods of transmission. As AIDS was studied, it became clear that the disease was spread through intimate contact such as sex and sharing syringes, as well as transfusions and other exposure to contaminated blood. It could not be spread through casual contact such as shaking hands, coughing, or sneezing. In 1984, Robert Gallo of the National Cancer Institute (NCI) and Luc Montagnier of the Institut Pasteur proved that AIDS was caused by a virus. There is still a controversy over priority of discovery. However, knowledge of the disease’s cause did not mean readiness to combat the plague. Homophobia and racism, combined with nationalism and fears of creating panic in the blood supply market, contributed to deadly delays before action was taken by any government. The relatively low number of sufferers skyrocketed around the world and created an uncontrollable epidemic. Immunology comes of age The 1980s were a decade of worldwide interest in immunology, an interest unmatched since the development of the vaccine era at the beginning of the century. In 1980, the Nobel Prize in Physiology or Medicine went to three scientists for their work in elucidating the genetic basis of the immune system. Baruj Benacerraf at Harvard University (Cambridge, MA), George Snell of the Jackson Laboratory (Bar Harbor, ME), and Jean Dausset of the University of Paris explored the genetic basis of the immune response. Their work demonstrated that histocompatibility antigens (called H-factors or H-antigens) determined the interaction of the myriad cells responsible for an immunological response. Early efforts to study immunology were aimed at understanding the structure and function of the immune system, but some scientists looked to immunology to try to understand diseases that lacked a clear outside agent. In some diseases, some part of the body appears to be attacked not by an infectious agent but by the immune system. Physicians and researchers wondered if the immune system could cause, as well as defend against, disease. By the mid-1980s, it was clear that numerous diseases, including lupus and rheumatoid arthritis, were connected to an immune system malfunction. These were called “autoimmune diseases” because they were caused by a patient’s own immune system. Late in 1983, juvenile-onset diabetes was shown to be an autoimmune disease in which the body’s immune system attacks insulin-producing cells in the pancreas. Allergies were also linked to overreactions of the immune system. By 1984, researchers had discovered an important piece of the puzzle of immune system functioning. Professor Susumu Tonegawa and his colleagues discovered how the immune system recognizes “self” versus “not-self”—a key to immune system function. Tonegawa elucidated the complete structure of the cellular T cell receptor and the genetics governing its production. It was already known that T cells were the keystone of the entire immune system. Not only do they recognize self and not-self and so determine what the immune system will attack, they also regulate the production of B cells, which produce antibodies. Immunologists regarded this as a major breakthrough, in large part because the human immunodeficiency virus (HIV) that causes AIDS was known to attack T cells. The T cell response is also implicated in other autoimmune diseases and many cancers in which T cells fail to recognize not-self cells. The question remained, however, how the body could possibly contain enough genes to account for the bewildering number of immune responses. In 1987, for the third time in the decade, the Nobel Prize in Physiology or Medicine went to scientists working on the immune system. As Tonegawa had demonstrated in 1976, the immune system can produce an almost infinite number of responses, each of which is tailored to suit a specific invader. Tonegawa showed that rather than containing a vast array of genes for every possible pathogen, a few genetic elements reshuffled themselves. Thus a small amount of genetic information could account for many antibodies. The immune system relies on the interaction of numerous kinds of cells circulating throughout the body. Unfortunately, AIDS was known to target those very cells. There are two principal types of cells, B cells and T cells. T cells, sometimes called “helper” T cells, direct the production of B cells, an immune response targeted to a single type of biological or chemical invader. There are also “suppressor” cells to keep the immune response in check. In healthy individuals, helpers outnumber suppressors by about two to one. In immunocompromised individuals, however, the T cells are exceedingly low and, accordingly, the number of suppressors extremely high. This imbalance appears capable of shutting down the body’s immune response, leaving it vulnerable to infections a healthy body wards off with ease. Eventually, scientists understood the precise mechanism of this process. Even before 1983, when the viral cause of the disease was determined, the first diagnostic tests were developed to detect antibodies related to the disease. Initially, because people at risk for AIDS were statistically associated with hepatitis, scientists used the hepatitis core antibody test to identify people with hepatitis, and therefore, at risk for AIDS. By 1985, a diagnostic method was specifically designed to detect antibodies produced against the low titer HIV itself. Diagnosing infected individuals and protecting the valuable world blood supply spurred the diagnosis effort. By the late 1980s, under the impetus and fear associated with AIDS, both immunology and virology received huge increases in research funding, especially from the U.S. government. Pharmaceutical technology and AIDS Knowing the cause of a disease and how to diagnose it is quite a different matter from knowing how to cure it. Although by no means the solution, ironically, the pharmaceutical technology initially used to combat HIV was discovered some 20 years before AIDS appeared. Azidothymidine (AZT) was developed in 1964 as an anticancer drug by Jerome Horowitz of the Michigan Cancer Foundation (Detroit). But because AZT was ineffective against cancer, Horowitz never filed a patent. In 1987, the ultimate approval of AZT as an antiviral treatment for AIDS was the result of both the hard technology of the laboratory and the soft technologies of personal outrage and determination (and deft use of the 1983 Orphan Drug Act). Initially, the discovery of the viral nature of AIDS resulted in little, if any, R&D in corporate circles. The number of infected people was considered too small to justify the cost of new drug development, and most scientists thought retroviruses were untreatable. However, Sam Broder, a physician and researcher at the NCI, thought differently. As clinical director of the NCI’s 1984 Special Task Force on AIDS, Broder was determined to do something. Needing a larger lab, Broder went to the pharmaceutical industry for support. As Broder canvassed the drug industry, he promised to test potentially useful compounds in NCI labs if the companies would commit to develop and market drugs that showed potential. One of the companies he approached was Burroughs Wellcome, the American subsidiary of the British firm Wellcome PLC. Burroughs Wellcome had worked on nucleoside analogues, a class of antivirals that Broder thought might work against HIV. Burroughs Wellcome had successfully brought to market an antiherpes drug, acyclovir. Although many companies were reluctant to work on viral agents because of the health hazards to researchers, Broder persevered. Finally, Burroughs Wellcome and 50 other companies began to ship chemicals to the NCI labs for testing. Each sample was coded with a letter to protect its identity and to protect each company’s rights to its compounds. In early 1985, Broder and his team found a compound that appeared to block the spread of HIV in vitro. Coded Sample S, it was AZT sent by Burroughs Wellcome. There is a long road between in vitro efficacy and shipping a drug to pharmacies—a road dominated by the laborious approval process of the FDA. The agency’s mission is to keep dangerous drugs away from the American public, and after the thalidomide scare of the late 1950s and early 1960s, the FDA clung tenaciously to its policies of caution and stringent testing. However, with the advent of AIDS, many people began to question that caution. AZT was risky. It could be toxic to bone marrow and cause other less drastic side effects such as sleeplessness, headaches, nausea, and muscular pain. Even though the FDA advocated the right of patients to knowingly take experimental drugs, it was extremely reluctant to approve AZT. Calls were heard to reform or liberalize the approval process, and a report issued by the General Accounting Office (GAO) claimed that of 209 drugs approved between 1976 and 1985, 102 had caused serious side effects, giving lie to the apparent myth that FDA approval automatically safeguarded the public. The agendas of patient advocates, ideological conservatives who opposed government “intrusion,” and the pharmaceutical industry converged in opposition to the FDA’s caution. But AIDS attracted more than its share of false cures, and the FDA rightly kept such things out of the medical mainstream. Nonetheless, intense public demand (including protests by AIDS activist groups such as act up) and unusually speedy testing brought the drug to the public by the late 1980s. AZT was hard to tolerate and, despite misapprehensions of its critics, it was never thought to be a magic bullet that would cure AIDS. It was a desperate measure in a desperate time that at best slowed the course of the disease. AZT was, however, the first of what came to be a major new class of antiviral drugs. Its approval process also had ramifications. The case of AZT became the tip of the iceberg in a new world where consumer pressures on the FDA, especially from disease advocacy groups and their congressional supporters, would lead to more and rapid approval of experimental drugs for certain conditions. It was a controversial change that in the 1990s would create more interest in “alternative medicine,” nutritional supplements, fast-track drugs, and attempts to further weaken the FDA’s role in the name of both consumer rights and free enterprise. Computers and pharmaceuticals In the quest for new and better drugs, genetic and biological technologies came to the fore in the 1980s. These included a combination of “hard” (machine-based) and “wet” (wet chemistry-based) technologies. In the early 1980s, pharmaceutical companies hoped that developments in genetic engineering and other fields of molecular biology would lead to sophisticated drugs with complicated structures that could act in ways as complicated and precise as proteins. Through understanding the three-dimensional structure and hence the function of proteins, drug designers interested in genetic engineering hoped they could create protein-based drugs that replicated these structures and functions. Unfortunately for the industry, its shareholders, and sick people who might have been helped by these elegantly tailored compounds, it did not turn out that way. By the end of the decade, it was clear that whatever economic usefulness there was in molecular biology developments (via increased efficiency of chemical drugs), such developments did not yet enable the manufacturing of complex biologically derived drugs. So while knowledge of protein structure has supplied useful models for cell receptors—such as the CD4 receptors on T cells, to which drugs might bind—it did not produce genetic “wonder drugs.” Concomitant with this interest in molecular biology and genetic engineering was the development of a new way of conceiving drug R&D: a new soft technology of innovation. Throughout the history of the pharmaceutical industry, discovering new pharmacologically active compounds depended on a “try and see” empirical approach. Chemicals were tested for efficacy in what was called “random screening” or “chemical roulette,” names that testify to the haphazard and chancy nature of this drug discovery approach. The rise of molecular biology, with its promise of genetic engineering, fostered a new way of looking at drug design. Instead of an empirical “try and see” method, pharmaceutical designers began to compare knowledge of human physiology and the causes of medical disorders with knowledge of drugs and their methods of physiological action to conceptualize the right molecules. This ideal design is then handed over to research chemists in the laboratory, who search for a close match. In this quest, the biological model of cell receptor and biologically active molecule serve as a guide. The role of hard technologies in biologically based drug research cannot be overstated. In fact, important drug discoveries owe a considerable amount to concomitant technological developments, particularly in imaging technology and computers. X-ray crystallography, scanning electron microscopy, NMR, and laser and magnetic- and optical-based imaging techniques allow the visualization of atoms within a molecule. This capability is crucial, as it is precisely this three-dimensional arrangement that gives matter its chemically and biologically useful qualities. Computers were of particular use in this brave new world of drug design, and their power and capabilities increased dramatically during the 1980s. The increased computational power of computers enabled researchers to work through the complex mathematics that describe the molecular structure of idealized drugs. Drug designers, in turn, could use the increasingly powerful imaging capabilities of computers to convert mathematical protein models into three-dimensional images. Gone were the days when modelers used sticks, balls, and wire to create models of limited scale and complexity. In the 1980s, they used computers to transform mathematical equations into interactive, virtual pictures of elegant new models made up of thousands of different atoms. Since the 1970s, the pharmaceutical industry had been using computers to design drugs to match receptors, and it was one of the first industries to harness the steadily increasing power of computers to design molecules. Software applications to simulate drugs first became popular in the 1980s, as did genetics-based algorithms and fuzzy logic. Research on genetic algorithms began in the 1970s and continues today. Although not popular until the early 1990s, genetic algorithms allow drug designers to “evolve” a best fit to a target sequence through successive generations until a fit or solution is found. Algorithms have been used to predict physiological properties and bioactivity. Fuzzy logic, which formalizes imprecise concepts by defining degrees of truth or falsehood, has proven useful in modeling pharmacological action, protein structure, and receptors. The business of biotechnology The best drug in the world is worthless if it cannot be developed, marketed, or manufactured. By the mid-1980s, small biotechnology firms were struggling for survival, which led to the formation of mutually beneficial partnerships with large pharmaceutical companies and a host of corporate buyouts of the smaller firms by big pharma. Eli Lilly was one of the big pharma companies that formed early partnerships with smaller biotech firms. Beginning in the 1970s, Lilly was one of the first drug companies to enter into biotechnology research. By the mid-1980s, Lilly had already put two biotechnology-based drugs into production: insulin and human growth hormone. Lilly produced human insulin through recombinant DNA techniques and marketed it, beginning in 1982, as Humulin. The human genes responsible for producing insulin were grafted into bacterial cells through a technique first developed in the production of interferon in the 1970s. Insulin, produced by the bacteria, was then purified using monoclonal antibodies. Diabetics no longer had to take insulin isolated from pigs. By 1987, Lilly ranked second among all institutions (including universities) and first among companies (including both large drug manufacturers and small biotechnology companies) in U.S. patents for genetically engineered drugs. By the late 1980s, Lilly recognized the link between genetics, modeling, and computational power and, already well invested in computer hardware, the company moved to install a supercomputer. In 1988, typical of many of the big pharma companies, Lilly formed a partnership with a small biotechnology company: Agouron Pharmaceuticals, a company that specialized in three-dimensional computerized drug design. The partnership gave Lilly expertise in an area it was already interested in, as well as manufacturing and marketing rights to new products, while Agouron gained a stable source of funding. Such partnerships united small firms that had narrow but potentially lucrative specializations with larger companies that already had development and marketing structures in place. Other partnerships between small and large firms allowed large drug companies to “catch up” in a part of the industry in which they were not strongly represented. This commercialization of drug discovery allowed companies to apply the results of biotechnology and genetic engineering on an increasing scale in the late 1980s, a process that continues. The rise of drug resistance New developments in the West in the late 1980s had particular implications for drugs and pharmaceutical technologies. Diseases that were thought to have been eliminated in developed countries reappeared in the late 1980s with a frightening twist: they had developed drug resistance. Tuberculosis in particular experienced a resurgence. In the mid-1980s, the worldwide decline in tuberculosis cases leveled off and then began to rise. New cases of tuberculosis were highest in less developed countries, and immigration was blamed for the increased number of cases in developed countries. (However, throughout the 20th century in the United States, tuberculosis was a constant and continued health problem for senior citizens, Native Americans, and the urban and rural poor. In 1981, an estimated 1 billion people—one-third of the world’s population—were infected. So thinking about tuberculosis in terms of a returned epidemic obscures the unabated high incidence of the disease worldwide over the preceding decades.) The most troubling aspect of the “reappearance” of tuberculosis was its resistance to not just one or two drugs, but to multiple drugs. Multidrug resistance stemmed from several sources. Every use of an antibiotic against a microorganism is an incidence of natural selection in action—an evolutionary version of the Red Queen’s Race, when you have to run as fast as you can just to stay in place. Using an antibiotic kills susceptible organisms. Yet mutant organisms are present in every population. If even a single pathogenic organism survives, it can multiply freely. Agricultural and medical practices have contributed to resistant strains of various organisms. In agriculture, animal feed is regularly and widely supplemented with antibiotics in subtherapeutic doses. In medical practice, there has been widespread indiscriminate and inappropriate use of antibiotics to the degree that hospitals have become reservoirs of resistant organisms. Some tuberculosis patients unwittingly fostered multidrug-resistant tuberculosis strains by failing to comply with the admittedly lengthy, but required, drug treatment regimen. In this context, developing new and presumably more powerful drugs and technologies became even more important. One such technology was combinatorial chemistry, a nascent science at the end of the 1980s. Combinatorial chemistry sought to find new drugs by, in essence, mixing and matching appropriate starter compounds and reagents and then assessing them for suitability. Computers were increasingly important as the approach matured. Specialized software was developed that could not only select appropriate chemicals but also sort through the potentially awesome number of potential drugs. Prevention, the best cure Even as the pharmaceutical industry produced ever more sophisticated drugs, there was a new emphasis in health care: preventive medicine. While new drugs were being designed, medicine focused on preventing disease rather than simply trying to restore some facsimile of health after it had developed. The link between exercise, diet, and health dates back 4000 years or more. More recently, medical texts from the 18th and 19th centuries noted that active patients were healthier patients. In the early 1900s, eminent physician Sir William Osler characterized heart disease as rare. By the 1980s, in many Western countries some 30% of all deaths were attributed to heart disease, and for every two people who died from a heart attack, another three suffered one but survived. During this decade, scientists finally made a definitive connection between heart disease and diet, cholesterol, and exercise levels. So what prompted this change in perspective? In part, the answer lies with the spread of managed care. Managed care started in the United States before World War II, when Kaiser-Permanente got its start. Health maintenance organizations (HMOs), of which Kaiser was and is the archetypal representative and which were and are controversial, spread during the 1980s as part of an effort to contain rising medical costs. One way to keep costs down was to prevent people from becoming ill in the first place. But another reason for the growth of managed care had to do with a profound shift in the way that diseases, particularly diseases of lifestyle, were approached. Coronary heart disease has long been considered the emblematic disease of lifestyle. During the 1980s, a causal model of heart disease that had been around since the 1950s suddenly became the dominant, if not the only, paradigm for heart disease. This was the “risk factor approach.” This view of heart disease—its origins, its outcomes, its causes—is a set of unquestioned and unstated assumptions about how individuals contribute to disease. Drawing from population studies about the relationship between heart disease and individual diet, genetic heritage, and habits, as much as from biochemical and physiological causes of atherosclerosis, the risk factor approach tended to be a more holistic approach. Whereas the older view of disease prevention focused on identifying those who either did not know they were affected or who had not sought medical attention (for instance, tuberculosis patients earlier in the century), this new conceptual technology aimed to identify individuals who were at risk for a disease. According to the logic of this approach, everyone was potentially at risk for something, which provided a rationale for population screening. It was a new way to understand individual responsibility for and contribution to disease. Focusing on risk factors reflected a cultural preoccupation with the individual and the notion that individuals were responsible for their own health and illness. As part of a wave of new conservatism against the earlier paternalism of the Great Society, accepting the risk factor approach implied that social contributions to disease, such as the machinations of the tobacco and food industries, poverty, and work-related sedentary lifestyles, were not to blame for illness. Individuals were considered responsible for choices and behavior that ran counter to known risk factors. While heart disease became a key focus of the 1980s, other disorders, including breast, prostate, and colon cancer, were ultimately subsumed by this risk factor approach. AIDS, the ultimate risk factor disease, became the epitome of this approach. Debates raged about the disease and related issues of homosexuality, condoms, abstinence, and needle exchange and tore at the fabric of society as the 1990s dawned. Conclusion The 1980s saw the resurgence of old perils, the emergence of new ones, and the rapid mobilization of new biotechnological tools to combat both. The decade also saw the surge and temporary fallback of entrepreneurial biotechnology. Throughout the decade, the hard technologies of genetic engineering and developments in immunology, automation, genomics, and computers—combined with the soft technologies of personal action and acknowledgment of risk and changes in government regulations and ways of thinking—affected the direction of biomedicine and pharmacy. The coming decade would see the explosive growth of these various trends—into both flowers and weeds. Harnessing genes, recasting flesh (1990s) In the 1990s, the Human Genome Project took off like a rocket, along with many global genomic initiatives aimed at plants, animals, and microorganisms. Gene therapy developed, as did the potential of human cloning and the use of human tissues as medicine. The new technologies provided hope for solving the failures of the old through a new paradigm—one that was more complex, holistic, and individualized. The changing view became one of medicines and therapies based on knowledge, not trial and error; on human flesh, not nature’s pharmacy. The new therapeutic paradigm evolved from an earlier promise of hormone drugs and the flowering of intelligent drug design. It grew from an understanding of receptors and from breakthroughs in genomics and genetic engineering. It found favor in the new power of combinatorial chemistry and the modeling and data management abilities of the fastest computers. Hope was renewed for a next generation of magic bullets born of human genes. As the 1990s ended, many genetically based drugs were in clinical trials and a wealth of genome sequences, their promise as yet unknown, led scientists to reason that the 21st century would be the biotech century. Computers and combinatorial technology Robotics and automation allowed researchers to finally break through a host of constraints on rational drug design. Achievements in miniaturization in robotics and computer systems allowed the manipulation of many thousands of samples and reactions in the time and space where previously only a few could be performed. They permitted the final transformation of pharmacology from the tedious, hit-and-miss science based primarily on organic synthesis to one based firmly on physiology and complex biochemistry, allowing explosive movement into rational drug discovery in both laboratory design and natural-product surveys. (And even when the technology remained hit-and-miss because of the lack of a compatible knowledge base, the sheer quantity of samples and reactions now testable created efficiencies of scale that made the random nature of the process extraordinarily worthwhile.) Combinatorial chemists produce libraries of chemicals based on the controlled and sequential modification of generally immobilized or otherwise tagged chemical starting blocks. These original moieties are chosen, under optimal knowledge conditions, for their predicted or possible behavior in a functional drug, protein, polymer, or pesticide. Developing the knowledge base for starting materials proved to be one of the greatest benefits of computer modeling of receptors and the development of computational libraries (multiple structures derived from computer algorithms that analyze and predict potentially useful sequences from databases of gene or protein sequences, or structural information from previous drugs). Here the search for natural products remained critical—for the discovery of new starting places. Finding new drug starting places, as much as finding drugs that were useful “as is,” became important in the 1990s as companies tried to tap into the knowledge base of traditional medical practitioners in the developing world through collaboration rather than appropriation. In this manner, several companies promoted the analysis of the biodiversity available in tropical rainforests and oceans. This “added value” of biodiversity became both the rallying cry for environmentalists and a point of solidarity for political muscle-building in the developing world. As the industrialized world’s demand for new drugs (and associated profits) increased, developing nations sought to prevent the perceived exploitation of their heritage. And just as previous sources of drugs from the natural world were not wholly ignored (even as the human genome became the “grail” of modern medical hopes), combinatorial approaches created a new demand for the services of traditional organic and analytical chemistry. Although computer models were beneficial, new wet chemistry techniques still had to be defined on the basis of the new discoveries in genomics and proteomics. They had to be modified for microsystems and mass production—for the triumph of the microtiter plate over the flask, the test tube, and the beaker. Drugs were still chemical products after all. High-throughput screening The vast increase in the number of potential drugs produced through combinatorial methods created a new bottleneck in the system—screening and evaluating these libraries, which held hundreds of thousands of candidates. To conduct high-throughput screening (HTS) on this excess of riches, new and more automated assays were pressed into service. Part of this move into HTS was due to the burgeoning of useful bioassays, which permitted screening by individual cells and cell components, tissues, engineered receptor proteins, nucleic acid sequences, and immunologicals. New forms of assays proliferated, and so did opportunities for evaluating these assays quickly through new technologies. Researchers continued to move away from radioactivity in bioassays, automated sequencing, and synthesis by using various new tagging molecules to directly monitor cells, substrates, and reaction products by fluorescence or phosphorescence. Fluorescence made it possible to examine, for the first time, the behavior of single molecules or molecular species in in vivo systems. Roger Tsien and colleagues, for example, constructed variants of the standard green fluorescent protein for use in a calmodulin-based chimera to create a genetic-based, fluorescent indicator of Ca2+. They called this marker “yellow chameleon” and used it in transgenic Caenorhabditis elegans to follow calcium metabolism during muscle contraction in living organisms. Of particular promise was the use of such “light” technologies with DNA microarrays, which allowed quantitative analysis and comparison of gene expression by multicolor spectral imaging. Many genes are differentiated in their levels of expression, especially in cancerous versus normal cells, and microarray techniques showed promise in the discovery of those differentiated genes. Microarrays thus became a basic research tool and a highly promising candidate for HTS in drug development. Genomics meets proteomics Knowledge of details of the genetic code, first learned during the 1960s, achieved practical application during the 1990s on a previously unimaginable scale. Moving from gene to protein and back again provided an explosion of information as the human (and other) genome projects racked up spectacular successes. Planned at the end of the 1980s, the U.S. Human Genome Project and the world Human Genome Organization led the way. Started first as a collection of government and university collaborations, the search for the human genome was rapidly adopted by segments of industry. The issue of patenting human, plant, and animal genes would be a persistent controversy. Inspired by this new obsession with genomics, the 1990s may ultimately be best known for the production of the first complete genetic maps. The first full microorganism genome was sequenced in 1995 (Haemophilus influenza, by Craig Venter and colleagues at The Institute for Genomic Research). This achievement was followed rapidly by the genome sequencing of Saccharomyces cerevisiae (baker’s yeast) in 1996; Escherichia coli, Borrelia burgdorferi, and Heliobacter pylori in 1997; the nematode C. elegans in 1998; and the first sequenced human chromosome (22) in 1999. The entrance of industry into the race to sequence the human genome at the tail end of the decade sped up the worldwide effort, albeit amid controversy. Hot on the heels of the genomics “blastoff” was the development of proteomics—the science of analyzing, predicting, and using the proteins produced from the genes and from the cellular processing performed on these macromolecules before they achieve full functionality in cells. Both proteomics and genomics rely on bioinformatics to be useful. Bioinformatics is essentially the computerized storage and analysis of biological data, from standard gene sequence databases (such as the online repository GenBank maintained by the NIH) to complex fuzzy logic systems such as GRAIL (developed in 1991 by Edward Eberbacher of Oak Ridge National Laboratory). GRAIL and more than a dozen other programs were used to find prospective genes in genomic databases such as GenBank by employing various pattern recognition techniques. Pattern recognition techniques were also at the heart of the new DNA microarrays discussed above, and they were increasingly used to detect comparative patterns of gene transcription in cells under various conditions and states (including diseased vs healthy). Human biotechnology In September 1990, the first human gene therapy was started by W. French Anderson at NIH in an attempt to cure adenosine deaminase (ADA) deficiency—referred to as bubble-boy syndrome—by inserting the correct gene for ADA into an afflicted four-year-old girl. Although the treatment did not provide a complete cure, it did allow the young patient to live a more normal life with supplemental ADA injections. Other attempts at gene therapy also remained more promising than successful. Clinical trials on humans were disappointing compared with the phenomenal successes in mice, although limited tumor suppression did occur in some cancers, and there were promising reports on the treatment of hemophilia. Jesse Gelsinger, a teenager who suffered from the life-threatening liver disorder ornithine transcarbamylase deficiency, volunteered for adenovirus-delivered gene therapy at a University of Pennsylvania clinical trial in 1999. His subsequent death sent a shock wave through the entire research community, exposed apparent flaws in regulatory protocols and compliance, and increased public distrust of one more aspect of harnessing genes. Using gene products as drugs, however, was a different story. From recombinant human insulin sold in the 1980s to the humanized antibodies of the 1990s, the successes of harnessing the human genome—whether “sensibly” (in the case of the gene product) or by using antisense techniques as inhibitors of human genes (in 1998 Formivirsen, used to treat cytomegalovirus, became the first approved antisense therapeutic)—proved tempting to the research laboratories of most major pharmaceutical companies. Many biotechnology medicines—from erythropoietin, tumor necrosis factor, dismutases, growth hormones, and interferons to interleukins and humanized monoclonal antibodies—entered clinical trials throughout the decade. Beginning in the 1990s, stem cell therapy held great promise. This treatment uses human cells to repair and ameliorate inborn or acquired medical conditions, from Parkinson’s disease and diabetes to traumatic spinal paralysis. By 1998, embryonic stem cells could be grown in vitro, which promised a wealth of new opportunities for this precious (and controversial) resource. Promising too were the new forms of tissue engineering for therapeutic purposes. Great strides were made in tissue, organ, and bone replacements. The demand for transplants, growing at 15% per year by the end of the decade, led to the search for appropriate artificial or animal substitutes. Cartilage repair systems, such as Carticel by Genzyme Tissue Repair, became commonplace. Patients’ cells shipped to the company were treated with Carticel, cultured, and subsequently reimplanted. Second-generation products permitted autologous cells to be cultured on membranes, allowing tissue formation in vitro. Several companies focused on developing orthobiologics—proteins, such as growth factors, that stimulate the patient’s ability to regenerate tissues. Epicel, a graft made from autologous cells, was also developed by Genzyme Tissue Repair to replace the skin of burn victims with greater than 50% skin damage. In 1998, Organogenesis introduced the first FDA-approved, ready-to-order Apligraf human skin replacement, which was made of living human epidermal keratinocytes and dermal fibroblasts. Also undergoing research in 1999 was Vitrix soft tissue replacement—which was made of fibroblasts and collagen. By the end of the decade, artificial liver systems (which work outside the body) were developed as temporary blood cleansers providing detoxification and various digestion-related processes. In many cases, such treatments allowed the patient’s own liver to regenerate during the metabolic “rest.” Such uses of cells and tissues raised numerous ethical questions, which were galvanized in the media by the 1996 arrival of the clonal sheep Dolly. Reaction against the power of the new biotechnology was not restricted to fears of dehumanizing humanity through “xeroxing.” The possibility of routine xenotransplantation (using animal organs as replacements in humans) came to the fore with advances in immunology and genetic engineering that promised the ability to humanize animal tissues (specifically, those of pigs) in ways similar to the development of humanized antibodies in mice. The issue of xenotransplantation not only raised fears of new diseases creeping into the human population from animal donors, but was seen by some as a further degradation of human dignity either intrinsically or through the misuse of animals. The animal rights lobby throughout the decade argued passionately against the use of animals for human health purposes. The Red Queen’s race Continuing the problems seen in the 1980s, old and new diseases were increasingly immune to the array of weapons devised against them. Like Alice in Through the Looking Glass, drug researchers had to run as fast as they could just to keep up in the race against bacterial resistance to traditional antibiotics (see Chapter 7). As the decade progressed, more diseases became untreatable with the standard suite of drugs. Various streptococcal infections, strains of tuberculosis bacteria, pathogenic E. coli, gonorrhea, and the so-called flesh-eating bacteria—necrotizing fasciitis, most commonly caused by group A streptococcus—all became immune to previously successful magic bullets. Patients died who earlier would have lived. As the problem manifested, pharmaceutical, software, and instrument companies alike turned to combinatorial chemistry and HTS technologies in an effort to regain the racing edge. AIDS remained a profoundly disturbing example of the failure of technology to master a disease, despite the incredible advances in understanding its biology that took place in the 1990s. Vaccine efforts occupied much of the popular press, and genetically engineered viral fragments seemed to be the best hope. But the proliferation of viral strains erased hope for a single, easy form of vaccination. Resistance to AZT therapy increased, and even the new protease inhibitors and so-called drug cocktails developed in the 1990s proved to be only stopgap measures as viral strains appeared that were resistant to everything thrown at them. Individual lives were prolonged, and the death rate in Western countries, where the expensive new drugs were available, dropped precipitously. But the ultimate solution to AIDS had not been found, nor even a countermeasure to its spread in the developing world and among poor populations of industrialized nations. The middle of the decade saw a resurgence of the “old” plague (bubonic) in India, and even polio remained a problem in the developing world. In Africa, Ebola resurfaced in scattered outbreaks—although it was nothing compared with the continental devastation caused by AIDS. In the rest of the world, fears of biological warfare raised by the Gulf War continued. Vaccination efforts were stepped up for many diseases. The Assembly of the World Health Organization set a global goal in 1990 of a 95% reduction in measles deaths in 1995 compared with pre-immunization levels. By the deadline, estimated global coverage for measles vaccine had reached 78%, at the same time as the industrialized world experienced a backlash against vaccination because of concerns about adverse side effects. Vaccine technology continued to improve with the development of recombinant vaccines for several diseases, new efforts to produce vaccines for old scourges such as malaria, and new nasal delivery systems that stimulated the mucosal-associated antibody system. DNA vaccines—the injection of engineered plasmids into human cells to stimulate antigen production and immunization—were first described in a 1989 patent and published in 1990 by Wolff, Malone, Felgner, and colleagues. They entered clinical trials in 1995. Although one editor of Bio/Technology called this the Third Vaccine Revolution, by the end of the decade the reality of this expansive claim remained in doubt, especially because of the continuing debate over genetic engineering. Efforts to develop food-based vaccines through the production of transgenics engineered with the appropriate antigens continued. This research stimulated studies on mucosal immunity and efforts to enable complex proteins to cross the gut–blood barrier. Even with these technological developments, the decade ended with the negatives of ever-expanding disease problems, exacerbated by predictions that global warming would lead to new epidemics of insectborne and tropical diseases. However, a note of optimism remained that rational design and automated production technologies would ultimately be able to defeat these diseases. High tech and new mech To improve the front end of rational drug design and to foster the use and growth of knowledge bases in genomics and proteomics, many old and new technologies were adapted to pharmaceutical use in the 1990s. In the 1990s, the use of mass spectrometry (MS) for bioanalytical analysis underwent a renaissance, with improvements such as ultrahigh-performance ms using Fourier-transform ion cyclotron resonance (FT-ICR MS) and tandem-in-time (multidimensional) MS for biological macromolecules. Improved techniques such as peak-parking (reducing the column flow rate into the mass spectrometer the instant a peak is detected, to allow samples to be split and analyzed by multiple MS nearly simultaneously) added several dimensions that were previously impossible. These changes improved the ability to analyze the complex mixtures required in studies of cellular metabolism and gene regulation. Results from multidimensional runs were analyzed by increasingly sophisticated bioinformatics programs and used to improve their knowledge base. In combination with HPLC and various capillary electrophoretic systems, ms became part of a paradigm for pharmaceutical R&D as a valuable new approach for identifying drug targets and protein function. Equivalently, the development of multidimensional NMR techniques, especially those using more powerful instruments (e.g., 500-MHz NMR) opened the door to solving the structure of proteins and peptides in aqueous environments, as they exist in biological systems. The new NMR techniques allowed observations of the physical flexibility of proteins and the dynamics of their interactions with other molecules—a huge advantage in studies of a protein’s biochemical function, especially in receptors and their target molecules (including potential drugs). By viewing the computer-generated three-dimensional structure of the protein, which was made possible by the data gathered from these instruments, the way in which a ligand fits into a protein’s active site could be directly observed and studied for the first time. The three-dimensional structure provided information about biological function, including the catalysis of reaction and binding of molecules such as DNA, RNA, and other proteins. In drug design, ligand binding by a target protein was used to induce the ultimate effects of interest, such as cell growth or cell death. By using new technologies to study the structure of the target protein in a disease and learn about its active or ligand-binding sites, rational drug design sought to design inhibitors or activators that elicited a response. This correlation between structure and biological function (known as the structure–activity relationship, or SAR) became a fundamental underpinning of the revolution in bioinformatics. In the 1990s, the SAR was the basis by which genomics and proteomics were translated into pharmaceutical products. The “new” Big Pharma Ultimately, the biggest change in the pharmaceutical industry, enabled by the progression of technologies throughout the century and culminating in the 1990s, was the aforementioned transformation from a hit-and-miss approach to rational drug discovery in both laboratory design and natural-product surveys. A new business atmosphere, first seen in the 1980s and institutionalized in the 1990s, revealed itself. It was characterized by mergers and takeovers, and by a dramatic increase in the use of contract research organizations—not only for clinical development, but even for basic R&D. Big Pharma confronted a new business climate and new regulations, born in part from dealing with world market forces and protests by activists in developing countries. Marketing changed dramatically in the 1990s, partly because of a new consumerism. The Internet made possible the direct purchase of medicines by drug consumers and of raw materials by drug producers, transforming the nature of business. Direct-to-consumer advertising proliferated on radio and TV because of new FDA regulations in 1997 that liberalized requirements for the presentation of risks of medications on electronic media compared with print. The phenomenal demand for nutritional supplements and so-called alternative medicines created both new opportunities and increased competition in the industry—which led to major scandals in vitamin price-fixing among some of the most respected, or at least some of the biggest, drug corporations. So powerful was the new consumer demand that it represented one of the few times in recent history that the burgeoning power of the FDA was thwarted when the agency attempted to control nutritional supplements as drugs. (The FDA retained the right to regulate them as foods.) Promise and peril At the start of the Pharmaceutical Century, the average life expectancy of Americans was 47. At century’s end, the average child born in the United States was projected to live to 76. As the 1900s gave way to the 2000s, biotechnology provided the promise of even more astounding advances in health and longevity. But concomitant with these technological changes was a sea change in the vision of what it was to be a human being. In the 19th century, the natural world was the source of most medicines. With the dawn of magic bullets in the 20th century, complex organic chemistry opened up a world of drugs created in the laboratory, either modified from nature or synthesized de novo. As the Pharmaceutical Century progressed, from the first knowledge of human hormones to the discovery of the nature of genes and the tools for genetic engineering, the modern paradigm saw a recasting of what human flesh was for—suddenly it was a source of medicines, tissues, and patentable knowledge. Humankind, not the natural world, became the hoped-for premier source of drug discovery. With genes and chemicals suddenly capable of manipulating the warp and woof of the human loom—both mind and body alike—the human pattern seemed to become fluid to design. According to pessimists, even if biotechnology were not abused to create “superhumans,” pharmaceuticals and health care could become the greatest differentiators of human groups in history—not by genetic races, but by economic factors. The new knowledge was found in the highest and most expensive technologies, and often in the hands of those more interested in patents than panaceas. Yet with this peril of inequality comes the promise of human transformation for good. According to optimists, biotechnology—especially the use of transgenic plants and animals for the production of new drugs and vaccines, xenotransplantation, and the like—promises cheaper, more universal health care. Meanwhile, lifestyle drugs—pharmaceuticals for nonacute conditions such as sterility, impotence, and baldness—also have emerged as a fast-growing category. From aspirin to Herceptin—a monoclonal antibody that blocks the overexpressed Her2 receptor in breast cancer patients—from herbal medicines to transgenic plants, from horse serum to xenotransplantation, from animal insulin to recombinant human growth hormone, the Pharmaceutical Century was one of transformation. It is too soon to predict what pattern the process will weave, but the human loom has gone high tech. The 21st century will be a brave new tapestry.