Book Read Free

Creating the Twentieth Century

Page 34

by Vaclav Smil


  Watt’s experiments and insights turned an extremely inefficient Newcomen’s machine with a very limited range of applications into the first widely useful mechanical prime mover energized by fuel combustion—but his 1769 patent, and its subsequent 25-year extension, actually impeded the next step of steam-driven innovation. Because Watt was afraid to work with high-pressure steam, he not only made no attempts to develop steam-driven transportation but also actively discouraged William Murdock, the principal erector of his engines, from doing so (Robinson and Musson 1969). Watt’s patent expired in 1800 and just a few years later Richard Trevithick and Oliver Evans had high-pressure boilers ready, and the diffusion of railways, steamships, and higher efficiency stationary engines was on the way. Some 80 years later, Karl Benz displayed a classic Watt syndrome by refusing to have anything to do either with very fast-running engines or vehicles other than motorized horse carriages.

  Marconi’s reluctance to broadcast anything else but Morse signals let others to take the lead in developing radio. And I have also noted, respectively, in chapters 2 and 1, Edison’s two famous failures of imagination, his militant rejection of AC and his belief that electricity, rather than internal combustion engines, will be the dominant automotive prime mover. In the first instance, Edison reversed himself fairly fast, but the second conviction gripped him for most of the first decade of the 20th century. Instead of trying to invent, for example, a reliable, and lucrative, electric starter for internal combustion engines (the device that was eventually designed in 1911 by Kettering), he persevered for years in his futile quest for a superior battery.

  Another display of failed imagination is a surprisingly common lack of confidence regarding the potential importance of one’s own inventions. Combative confidence and exaggerated expectations might be expected as a part of aggressive inventive process—but not an almost inexplicable diffidence and technical timidity. Among the most famous examples are Elisha Gray’s initial decision not to pursue his rights to patenting telephone because he believed that the device is just an interesting toy, Hertz’s complaints to his students that the electromagnetic oscillations he just discovered cannot possibly have any practical use (think of today’s universe of electronic devices!), and Louis Lumieère’s initial conviction that the cinema is an invention without any future as he expected that people will get bored watching his images of city scenes, trains, workers, and landscapes, most of which they could see just by walking around.

  Besides the surprising failures of imagination, there were also some notable personal tragedies. While most of the era’s inventors had fulfilling careers and led interesting lives, there were also suicides, premature deaths, and, more frequently, convictions of being unjustly treated. I have described Diesel’s feelings of failure and his (almost certain) suicide (chapter 3). Ironically, by the time of Diesel’s disappearance in 1913, his great invention was well set on its road to a commercial triumph, while an earlier sudden death may have changed the history of the world’s most popular entertainment. In September 1890, Louis Le Prince boarded a train in Dijon, after a visit to his brother, and never made it to Paris. We will never know why, but we can ask: if Le Prince had lived for at least another five years, could he had perfected the camera with which he took in 1888 what are arguably the world’s first moving pictures, and would he be known today as the inventor of cinema?

  Alfred Nobel did not commit suicide and did not die a mysterious death, but one of his early experiments with nitroglycerine killed his youngest brother, Emil, and four other people, and his near-chronic depression, despondency, and searing self-appraisal would have made Freud shudder. This is how, at age 54, when refusing his brother’s request to write an autobiography, Nobel saw himself (cited in Fant 1993:1):

  Alfred Nobel—a pitiful creature, ought to have been suffocated by a humane physician when he made his howling entrance into his life. Greatest virtues: keeping his nails clean and never being a burden to anyone. Greatest weaknesses: having neither wife and kids nor sunny disposition nor hearty appetite. Greatest single request: to not be buried alive. Greatest sins: not worshiping Mammon. Important events in his life: none.

  Two years later, in a letter to Sofie Hess, an Austrian flowers salesgirl whom he met as a 20-year-old in 1876 and who was his mistress for the next 18 years, he wrote (quoted in Fant 1993:318): “[W]hat a sad end I am going toward, with only an old servant who asks himself the whole time if he will inherit anything from me. He cannot know that I am not leaving a last will…”But, of course, he did, and probably the world’s most famous one: it divided his diminished but still considerable riches among 19 relatives, coworkers, and acquaintances (20% of the total), various institutions (16%), and the Swedish Academy of Sciences in order to “constitute a fund, the interest on which shall be annually distributed in the form of prizes to those who, during the preceding year, shall have conferred the greatest benefit on mankind” (Nobel 1895).

  While aging Nobel was depressed but lucid, the last two decades of Tesla’s life were marked by a deepening psychosis. He told journalists of his great love for a white pigeon (“Yes, I loved her as man loves a woman …”), and, finally a few days before his death on January 7, 1943, he sent a messenger with a sealed envelope containing money and addressed to Samuel Clemens at 35 South Fifth Avenue. That was the address of Tesla’s first New York laboratory of the late 1880s, and in 1943 Mark Twain, whom Tesla wanted to help (“He was in my room last night … He is having financial difficulties …”) had been dead for 33 years (Cheney 1981).

  Mergenthaler found he had tuberculosis even before he was 40; he moved from Baltimore to New Mexico, but there his house, and all of his papers, were destroyed by fire, and he died in Baltimore before his 46th birthday (Kahan 2000). Hertz died at the age of 36, after he became ill with an infection of the mouth and ears and underwent several unsuccessful surgeries. Wilbur Wright died of typhoid at age 45. Daimler’s long and accomplished life had an ironic ending. He did not like to drive (and may have actually never driven at all), and his death came after he insisted, with his health failing, on being taken in poor weather in an open vehicle to inspect a possible site for a new factory. On the return trip he collapsed, fell out of the car, and died soon afterward.

  But what beset his fellow inventor’s family was infinitely more tragic. As Maybach’s new powerful and elegant Mercedes 35 was breaking the world speed records in 1901 and 1902, his adolescent son Adolf (born in 1884) was succumbing to schizophrenia. His condition later deteriorated to such an extent that he could not be kept at home and had to be cared for in sanatoriums. Yet the worst was to come decades later: Adolf was murdered in 1940, 11 years after his father’s death, as one of the thousands of victims of the Nazi Euthanasieprogramm. Who remembers this as the company is now advertising the rebirth of a great marquee whose vehicles now retail more than $350,000?

  Not a few inventors discovered that even an extensive record of remarkable accomplishments did not automatically translate into widespread recognition and financial rewards. Fessenden felt slighted both in his native Canada and in the United States (Raby 1970). He got his first substantial reward only for his sonar patent, and received a large settlement for his pioneering radio inventions only in 1928, after years of litigation and just four years before his death. Much as today, willingness to engage in self-promotion and a dose of unorthodox behavior helped to capture press attention and to create an erroneous but greatly appealing image of a prototypical heroic inventor who appears to base his work on nothing but acute intuition and succeeds due to his uncommon perseverance.

  Edison was the master of this game, while Tesla was too eccentric to play it successfully. This difference helps to explain why Edison remains, more than 70 years after his death, widely admired by the public that craves such suitably heroic figures and why Tesla’s following—although highly, and some of it even fanatically, devoted to the memory of that master electrician—is much more cultlike, and why outside of his native Serbi
a (where he will always be a national hero) it is largely limited to scientists and to individuals who are intrigued by his research into ultrahigh-frequency discharges, large-scale energy transmissions without wires, and death rays. However, Tesla got one honor that eluded Edison: the unit of magnetic flux density carries his name, and so he joins the most select company of scientists and engineers—including Ampèere, Coulomb, Faraday, Hertz, Ohm, and Volta—whose names were chosen for international scientific units.

  But public admiration and official honors went not only to such skilled self-promoters as Edison or Marconi but also to inventors whose work is now known only to historians of science and engineering. In the words of his brother, William Siemens (see figure 4.5) “forced the public opinion of England to honour him in his lifetime, and in a still more striking manner after his death” (Siemens 1893:270). He was knighted, got honorary degrees from both Oxford and Cambridge, and on November 21, 1883, the London Times obituary spoke of his “singularly powerful and fertile mind.” His funeral service took place in Westminster Abbey, and later a window in the cathedral was dedicated to his memory: England’s highest honors for an immigrant German engineer.

  I end these reflections on personalities by noting a few idiosyncrasies that have been a surprisingly frequent accompaniment of creative process. Edison could not just nap but sleep deeply just about anywhere—fully dressed in a much crumpled three-piece suit lying on wooden desks and benches (figure 6.4) and on bare floors, much like an ascetic Chinese sage with only a bent arm for his pillow. During the construction of the world’s first electric network, his company kept a large stock of tubes in the cellar of the station at Pearl Street. Edison recalled that “as I was on all the time, I would take a nap of an hour or so in the daytime—and I used to sleep on those tubes in the cellar. I had two Germans who were testing there, and both of them died of diphtheria, caught in the cellar, which was cold and damp. It never affected me” (quoted in Dyer and Martin 1929:400).

  FIGURE 6.4. Thomas Edison snatching a nap on a wooden bench. This image is available from the National Park Service.

  Cherish the thought of perhaps the most influential inventor of a new era reposing on iron pipes in a dingy cellar—and try to imagine one of today’s CEOs getting involved in the same, pun unintended, down-to-earth fashion. In contrast, there were Tesla’s obsessive neuroses about germs and cleanliness: he eventually required even his closest friends to stand at a distance lest they contaminate him—although he did not worry at all about close contact with thousands of pigeons, including many in his hotel rooms, that he cared for throughout his life (Cheney 1981). The Wright brothers were unsurpassed workaholics (“two of the workingest boys I ever saw,” as one acquaintance put it) who had no friends outside the immediate family (Tobin 2003). In 1926, 14 years after Wilbur’s death, their sister Katharine, 52 years old at that time, decided to marry Henry Haskell. Orville thought this a betrayal of the family, and he not only refused to attend the wedding but also cut all contacts with her until shortly before her death in 1929.

  So Much Had Changed

  Byrn (1896:82) noted in his essay on technical progress that

  [i]t is so easy to lose sight of the wonderful, whence familiar with it, that we usually fail to give the full measure of positive appreciation to the great things of this great age. They burst upon our vision at first like flashing meteors: we marvel at them for a little while, and then we accept them as facts, which soon become so commonplace and so fused into the common life as to be only noticed by their omission.

  And what an omission that would be. The best way to appreciate the enormity of technical advances that took place between 1867 and 1914 is to try to construct a modern world devoid of just 10 major achievements introduced during that era.

  No electricity and hence no nonpolluting, convenient, and inexpensive lights and no electric motors to power myriads of precisely controllable machines, appliances, and trains; no internal combustion engines and hence no fast and affordable motorized vehicles, no freedom of personal movement, and also no airplanes and no effortless long-distance travel; no reproduction of sound and hence no telephones and music recordings; no photographic film and hence no convenient cameras and no movies; no way to generate, broadcast, and receive electromagnetic waves and hence no radio or TV; no steel alloys and no aluminum and hence no skyscrapers, no affordable machines and appliances; no easy way to produce inexpensive paper or to reproduce images and hence no mass publication of books and periodicals; and no nitrogen fertilizers and hence very widespread malnutrition, shortened life spans, and a world that could not feed its population.

  Byrn (1896:6) thought that this would leave “such an appalling void that we stop short, shrinking from the thought of what it would mean to modern civilization to eliminate from its life these potent factors of its existence.” The void would be so profound that what would exist would be only a prelude to modern civilization that was already in place by the middle of the 19th century: a society relying on steam engines and draft horses, whose nights would be sparsely and dimly lit by kerosene and coal gas, whose dominant metal would be brittle cast iron, and whose best means of long-distance communication would be a telegraphic message.

  And what was no less remarkable than this wealth of innovations was their lasting quality. Inevitably, not a few technical advances that were introduced during the Age of Synergy and that later came to dominate their respective markets were eventually replaced by new, superior 20th-century designs (e.g., internal combustion engines displaced by gas turbines in long-distance flights) or were relegated to small niche markets (the gramophone is still preferred by some audiophiles). But the opposite is true in a remarkably large number of cases where not only the basic design but even particular forms and modes of operation have endured with minimal amount of adjustments or even survived in virtually unchanged form.

  Enduring Artifacts and Ubiquitous Innovations

  Perhaps the most impressive way to illustrate enduring qualities of so many pre-WWI artifacts is a do-it-yourself exercise of comparing many objects and machines that we use today with the designs introduced, and rapidly improved, four to five generations ago. Most people will never have a chance to compare an 1880s Thomson dynamo or a large diesel engine built before 1910 with their modern counterparts. Consequently, such comparisons are done most conveniently with smaller objects of everyday use with which everybody is familiar. Light bulbs and spark plugs are excellent examples. Edison and Bosch could almost mistake one of today’s standard incandescent lights or spark plugs for one of their own designs.

  Modern light bulbs are machine-made rather than mouth-blown, their filaments are different, and Edison’s lamps were not filled with inert gases. But when hidden by frosting (which was applied externally to some early lamps since the 1880s, internally by GE as early as 1903), the only outward feature that would give an old lamp immediately away would be the glass tip surmounting the globe. The shape, size, and proportions of GE’s basic 100-W Soft White bulb made in 2000 are very similar to Edison’s lamps made during the 1890s (figure 6.5). A gently angled neck widens into a spherical top whose diameter accounts for just more than half of the light bulb’s overall length; a metal screw with four turns is another shared feature.

  More important, the lamp’s basic operational ratings are virtually identical to specifications set down by Edison and Upton in 1879: consuming 100 W of electricity at 120 V, a modern light draws the current of 0.83 A and has a resistance of 144.5Ω. And if we move to 1913 and compare that year’s tipless, internally frosted, gas-filled, coiled tungsten-filament lamp with today’s 100-W Soft White, we have two almost identical items. There may be no better example of a relatively complex artifact that has remained basically unchanged during nine decades of the most rapid technical innovation in history, and that had been produced in billions of copies to become one of the most common possessions of the 20th century.

  Spark plugs made a century apart share every key co
mponent: a terminal nut that connects to a spark plug wire, the metal-core center electrode that projects from the porcelain insulator nose, the ground electrode welded to the threaded part of the shell that forms the plug’s reach, and the hexagonal section in the upper part of the shell used to tighten the device by the spark plug wrench. One more example of an enduring design is decidedly low-tech (figure 6.6). In 1892 William Painter invented a clever way

  FIGURE 6.5. Comparison of two light bulbs made a century apart: GE’s standard frosted 100-W bulb made in the year 2000 (left) and Edison’s low-voltage carbon-filament lamp from the 1890s (its flat buttonlike contact is a key dating feature).

  for the sealing of bottles by the use of compressible packing disks and metallic caps, which have flanges bent into reliable locking engagement with annular locking shoulders on the heads of bottles, while the packing-disk is in each case under heavy compression and in enveloping contact with the lip of the bottle (Painter 1892:1).

  And crimped metal bottle caps are also an excellent example of innovations that are not examined in this book, which concentrates on fundamental, first-order advances and on their direct derivatives that often followed the primaries with an admirable speed. Electricity generation, internal combustion engines, inexpensive steels, aluminum, ammonia synthesized from its elements, and transmission of electromagnetic waves are the key primaries; electric motors, automobiles, airplanes, skyscrapers, recorded sound, and projected moving images are the ubiquitous derivatives. But there was much more to the period than introducing, rapidly improving, and commercializing those epoch-making inventions. There were many new inventions of simple objects of everyday use whose mass manufacturing was made possible, or commercially viable, thanks to the availability of new or better, and also cheaper, materials. Consequently, those inventions can be classed as secondary, or perhaps even tertiary, derivations.

 

‹ Prev