Book Read Free

Creating the Twentieth Century

Page 5

by Vaclav Smil


  What legacies of those two incomparable generations we will take with us into the 21st century? Closing decades of the 20th century had brought some unprecedented new technical capabilities, but by the year 2000 the New Economy was in retreat. I will not speculate what the post-1999 realities mean for the long term. But, averse as I am to engage in any long-range forecasting, I am on very solid ground when I keep pointing out repeated failures of enthusiastic forecasts that have been predicting the demise of some well-established techniques and imminent adoption of new technical and managerial approaches (Smil 2000a).

  Imminent demise of internal combustion engine is perhaps the best example in the first category. Edison believed that engines will never really make it and that the future belongs to electric cars—and he spent almost the entire first decade of the 20th century in a stubborn quest to develop a battery whose energy density would rival that of gasoline. A century later we still do not have an electricity storage device that would even approach such a density—but the promise of electric cars is still with us as forecasts that promise how these vehicles will capture a significant share of the automobile market within a decade are being rescheduled for another decade once the original forecast fizzles. A no less notorious example of forecasting excess includes all those monotonously advancing predictions of electricity from nuclear fusion that see the success always 30–50 years ahead.

  Most recent examples of unrealistic expectations include the forecasts of rapid diffusion of fuel cells that will usher a new age of hydrogen economy—in a clear disregard of the past experience, which shows that long lead times are necessary in order to accomplish such transitions (Smil 2003)—and the promise of paperless offices saving trees everywhere: in reality, the consumption of paper has been rising along with the diffusion of electronic word processing and data management. These and other similarly conspicuous forecasting failure only strengthen my conviction that many epoch-making innovations that were introduced before WWI and that had served us so well, and often with no fundamental modifications, during the 20th century have much greater staying power than may be commonly believed. We will continue to rely on them for much of the 21st century, and this realization makes it even more desirable, as well as more rewarding, to understand the genesis of these innovations.

  FIGURE 1.11 Grandiose, elegantly domed, and richly decorated central hall of Grands Magasins du Printemps (owned by Jules Jaluzot & Cie.) on Boulevard Haussman in Paris, the city’s premiere luxury department store. Its three wings and eight floors of diverse merchandise reflected the rise of affluent middle class during the two pre-WWI generations. Reproduced from The Illustrated London News, October 18, 1884.

  2

  The Age of Electricity

  There is a powerful agent, obedient, rapid, easy, which conforms to every use, and reigns supreme on board my vessel. Everything is done by means of it. It lights it, warms it, and is the soul of my mechanical apparatus. This agent is electricity.

  Captain Nemo to Pierre Aronnax in Jules Verne’s

  Twenty Thousand Leagues Under the Sea (1870)

  I do not know of any better description of electricity’s importance in modern society than taking this quotation from Jules Verne’s famous science fiction novel and substituting “in modern civilization” for “on board my vessel.” In 1870, when Verne set down his fictional account of Nemo’s global adventures, various electric phenomena had been under an increasingly intensive study for more than a century (Figuier 1888; Fleming 1911; MacLaren 1943; Dunsheath 1962). Little progress followed the pioneering 17th-century investigations by Robert Boyle (1627–1691) and Otto von Guericke (1602–1686). But Pieter van Musschenbroek’s (1692–1761) invention of the Leyden jar (a condenser of static electricity) and Benjamin Franklin’s (1706–1790) bold and thoughtful experiments in Boston (beginning in 1749) and later in Paris revived the interest in the properties of that mysterious force.

  FRONTISPIECE 2. Cross section of Edison’s New York station (thermal capacity, 93 MW) completed in 1902. Boilers are on the right; steam engine and dynamo hall on the left. Four steel-plate stacks were 60 m tall. Reproduced from the cover page of Scientific American, September 6, 1902.

  Electricity ceased to be a mere curiosity and became a subject of increasingly systematic research with the work by Luigi Galvani (1737–1798, of twitching frog legs fame, during the 1790s), Alessandro Volta (1745–1827; his pioneering paper that described the construction of the first battery was published in 1800), Hans Christian 0rsted (1777–1851, uncovered the magnetic effect of electric currents in 1819), and André Marie Ampere (1775–1836), who contributed the concept of a complete circuit and quantified the magnetic effects of electric current. Their research opened up many new experimental possibilities, and already in 1820 Michael Faraday (1791–1867), using 0rsted’s discovery, built a primitive electric motor, and before 1830 Joseph Henry’s (1797–1878) experimental electromagnets became powerful enough to lift briefly loads of as much as 1 t.

  But it was Michael Faraday’s discovery of the induction of electric current in a moving magnetic field in 1831 that eventually led to the large-scale conversion of mechanical energy into electricity (Faraday 1839). Its revelation is easily stated: the magnitude of the electromotive force that is induced in a circuit is proportional to the rate of change of flux. This discovery of how to generate alternating current (AC) was the logic gate that opened up the route toward practical use of electricity that was not dependent on bulky, massive, low-energy-density batteries. Eventually, this route had led to three classes of machines whose incessant, highly reliable, and remarkably efficient work makes it possible to permeate the modern world with inexpensive electricity: turbogenerators, transformers, and electric motors.

  Shortly after Faraday’s fundamental discoveries came the invention of telegraph that by the 1860s evolved into a well-established, globe-spanning way of wired communication. But by 1870 there were no commercially viable electric lights and, as the puzzled Aronnax noted when told about electric propulsion of the Nautilus by Nemo, until that time electricity’s “dynamic force has remained under restraint, and has only been able to produce a small amount of power”—because they were no reliable means of large-scale electricity generation and because electric motors were limited to small, battery-operated units.

  Experiments with electricity and the expanding telegraphy were energized by batteries, commonly known as Voltaic bimetallic piles whose low energy density (< 10 Wh/kg) was suited only for applications that required limited power. Then, after millennia of dependence on just three basic sources of energy—combustion of fuels (biomass or fossil), animate metabolism (human and animal muscles), and conversion of indirect solar flows (water and wind)—everything changed in the course of a single decade. During the 1880s the combined ingenuity of inventors, support of investors, and commercially viable designs of enterprising engineers coalesced into a new energy system without whose smooth functioning there would be no modern civilization.

  Practical carbon-filament electric lights, soon supplanted by incandescing finely drawn wires, illuminated nights. Parsons’s invention of the steam turbine created the world’s most powerful prime mover that had made the bulky and inefficient steam engines obsolete and allowed inexpensive generation of electricity on large scale. Transformers made it possible to transmit electricity over increasingly longer distances. Efficient induction motors, patented first in 1888 by Nikola Tesla, converted the flow of electrons into mechanical energy, and innovative electrochemistry began producing new materials at affordable prices. Economic and social transformations brought by electricity were so profound because no other kind of energy affords such convenience and flexibility, such an instant and effortless access to consumers: a flip of a switch, or even just a preprogrammed order.

  To any homemaker or any laborer of the pre-electrical era, electricity was a miracle. In households it had eventually eliminated a large number of daily chores that ranged from tiresome—drawing and
hauling water, washing and wringing clothes by hand, and ironing them with heavy wedges of hot metal—to light but relentlessly repetitive tasks (e.g., trimming wicks and filling oil lamps). Electricity on farms did away with primitive threshing of grains, hand-milking of cows, laborious preparation of feed by manual chopping or grinding, and pitchforking of hay into lofts. And, of course, in irrigated agricultures electric pumps eliminated slow and laborious water-raising techniques powered by people and animals. In workshops and factories, electricity replaced poorly lit premises first by incandescent and later by fluorescent lights while convenient, efficient, and precisely adjustable electric motors did away with dangerous transmission belts driven by steam engines. On railways, electricity supplanted inefficient, polluting steam engines.

  No other form of energy can equal electricity’s flexibility: it can be converted to light, heat, motion, and chemical potential, and hence it has been used extensively in every principal energy-consuming sector except commercial flying. Until 1998 the last sentence would have said simply “with the exception of flying”—but in that year AeroVironment’s unmanned Pathfinder aircraft rose to 24 km above the sea level. In August 2001, a bigger Helios, a thin, long curved, and narrow flying wing (its span of just over 74 m is longer than that of a Boeing 747) driven by 14 propellers powered by 1 kW of bifacial solar cells, became the world’s highest flying plane as it soared to almost 29 km (figure 2.1; AeroVironment 2004).

  Besides this all-encompassing versatility, electricity is also a perfectly clean, as well as a silent, source of energy at the point of consumption, and its delivery can be easily and precisely adjusted in order to provide desirable speed and accurate control for flexible industrial production (Nye 1990; Schurr et al. 1990). Electricity can be converted without virtually any losses to useful heat, and large electric motors can transform more than 90% of it into mechanical energy; it can also generate temperatures higher than combustion of any fossil fuel, and once a requisite wiring is in place any new converters can be just plugged in. Perhaps the best proof of electricity’s importance comes from simply asking what we would not have without it. The answer is just about everything in the modern world.

  FIGURE 2.1. Helios prototype flying wing, powered by photovoltaic cells, during its record-setting flight above the Hawaiian Islands on July 14, 2001. NASA photo ED 010209-6.

  We use electricity to power our lights, a universe of electronic devices (from cell phones to supercomputers), a panoply of converters ranging from handheld hair dryers to the world’s fastest trains, and almost every life saver (modern synthesis and production of pharmaceuticals is unthinkable without electricity, vaccines need refrigeration, hearts are checked by electrocardiograms and during operations are bypassed by electric pumps), and most of our food is produced, processed, distributed, and cooked with the help of electric machines and devices. This chapter’s goal is thus simple: to describe the genesis and evolution of electric systems that took place between the 1870s, the last pre-electric decade in history, and the first decade of the 20th century. During this span of less than two generations, we made an enormous progress as we put in place the foundation of a new energy system whose performance is now far ahead of anything we had before WWI but whose basic features remained remarkably constant throughout the 20th century.

  Fiat Lux : Inventing Electric Lights

  Thomas Edison was accustomed to keeping a brutal work pace and demanded others to follow, and his search for a durable filament that would produce more than an ephemeral glow was particularly frustrating (Josephson 1959; Israel 1998). But the often repeated dramatic story of continuous and sus-penseful “death-watch” that took 40 hours of waiting for the first successful filament to stop incandescing is just a legend, derived from later reminiscences of Edison’s assistant (Jehl 1937). Edison’s laboratory records indicate that the lamp that made history by passing the 10-hour mark—the one with a very fine carbonized cotton filament, a piece of six-cord thread fastened to platinum wires—was attached to an 18-cell battery at 1:30 A.M. on October 22, 1879, and that it still worked by 3 P.M. and continued to do so for another hour after increased power supply overheated and cracked the bulb (Friedel and Israel 1986).

  A filament incandescing for nearly 15 hours represented a major advance in the quest for electric light, but as far as materials and design procedures were concerned, it was a step clearly retracing previous research of other inventors. Edison’s principal, and lasting, contribution to the development of electric light was in changing the basic operating conditions, not in discovering new components. Even then, the initial success had to be followed by numerous improvements and modifications, in Edison’s own laboratory and by others, before reliable and reasonably long-lasting lightbulbs were ready for mass marketing. What is so remarkable is that so many basic material and operational components that were finessed during the 1880s are still very recognizably with us, and that many subsequent innovations of incandescent lighting have greatly improved its performance without changing their operating principles.

  Early Electric Lights

  The quest to use electricity for lighting began decades before Edison was born. The possibility to do so was first realized at the very beginning of the 19th century. In 1801 Humphry Davy (1778–1829) was the first scientist to describe the electric arc that arises as soon as two carbon electrodes are slightly separated. In 1808 Davy publicly demonstrated the phenomenon on a large scale at the Royal Institution in London by using 2,000 Voltaic cells to produce a 10-cm-long arc between the electrodes of willow charcoal (Davy 1840). First trials of arcs preceded by decades the introduction of large electricity generators. The world’s premiere of public lighting took place in December 1844 when Joseph Deleuil and Léon Foucault briefly lit Place de la Concorde by a powerful arc (Figuier 1888).

  The first dynamo-driven outdoor arc lights were installed during the late 1870s: starting in 1877, P. N. Yablochkov’s (1847–1894) much admired electric candles were used to illuminate downtown streets and other public places in Paris (Grand Magasins du Louvre, Avenue de l’Opeéra) and London (Thames Embankment, Holborn Viaduct, the British Museum’s reading room). By the mid-1880s arc lamps were fairly common sights in many Western cities, but they were massive and complicated devices that required skilled installation and frequent maintenance. Because a continuous arc wears away the electrodes, a mechanism was needed to move the rods in order to maintain a steady arc, and also to rejoin them and separate them once the current was switched off and on.

  Many kinds of self-regulating mechanisms were invented to operate arc lamps, and later designs did not need any complex regulators because they used several pairs of parallel upright carbons (separated by plaster), each able to operate for 90–120 minutes and switched on, manually or automatically, by bridging the two tips. But that still left considerable costs of the carbons and the labor for recarboning the devices. A typical 10-A arc lamp using, respectively, carbons of 15 and 9 mm for its positive and negative electrodes and operating every night from dusk to dawn would have consumed about 180 m of carbon electrodes a year (Garcke 1911a). Placing these 500-W lamps 50 m apart for basic street illumination would have required annual replacement of 3.6 km of carbons for every kilometer of road, a logistically cumbersome and costly proposition precluding the adoption of arc lamps for extensive lighting of public places.

  Glass-enclosed arcs extended the lamp’s useful life to as much as 200 hours, but they did not lower the cost of recarboning enough to make arcs a better choice than advanced gas lighting, and they also reduced the maximum luminosity, from as much as 18,000 to no more than 2,500 lumens (lm). As for the electricity supply, even a typical 10-A arc consuming 500–600 W had to be supplied by a line of just 50–60 V, a voltage too low for efficient large-scale electricity distribution. Obviously impractical for most indoor, and for all household, uses and uneconomical for the outdoor applications, arc lights could not become the sources of universal illumination. Less powerful and much more
practical incandescent lights were the obvious choice to fill this enormous niche—and it was also Humphry Davy who demonstrated their possibility when he placed a 45-cm-long piece of platinum wire (diameter of 0.8 mm) in a circuit between the bars of copper and induced first red heat and then brilliant white light (Davy 1840).

  A tedious account of inventors and their failed or promptly superseded attempts would be needed to review all of the activities that took place between 1820, when William de La Rue experimented with a platinum coil, and January 27, 1880, when Edison was granted his basic patent (U.S. Patent 223,898) for carbon-filament electric lamp (Edison 1880a; figure 2.2). More than a score of inventors in the United States, the United Kingdom, France, Germany, and Russia sought patents for the “subdivided” electric light for nearly four decades before Edison began his experiments (Pope 1894; Howell and Schroeder 1927; Bright 1949; Friedel and Israel 1986; Bowers 1998). In his February 1913 lecture at the New York Electrical Society, William J. Hammer, one of Edison’s laboratory assistants at Menlo Park during the period of lightbulb experiments and later the electrician of the first Edison lamp factory, pleaded “that we should keep these names green in our memory because their work was work of importance, even though it did not directly result in the establishment of the commercial incandescent electric lamp” (Hammer 1913:2).

 

‹ Prev