Nuclear Weapons

views updated May 23 2018

Nuclear Weapons

How nuclear weapons work

Effects of nuclear weapons

Nuclear weapons today

Resources

Nuclear weapons are explosive devices that release nuclear energy. An individual nuclear weapon may have an explosive force equivalent to millions of tons (megatons) of trinitrotoluene (TNT, the chemical explosive traditionally used for such comparisons) and can completely destroy a large city.

The destructive power of nuclear weapons derives from the core of the atom, the nucleus. One type of nuclear weapon, the fission bomb, uses the energy released when nuclei of heavy elements such as plutonium fission or split apart. Another, even more powerful type of nuclear weapon, the fusion or hydrogen bomb, uses the energy released when nuclei of hydrogen are fuse or unite.

Nuclear devices have been produced in many sizes and for many purposes. Bombs are devices that can be dropped from airplanes; warheads can be delivered by missiles launched from land, air, or sea, or by torpedoes; artillery shells can be fired from cannon; mines can be placed on the land or in the sea. Some nuclear weapons are small enough to destroy only a portion of a battlefield and are called tactical others, as already mentioned, are large enough to destroy entire cities or other major targets and are called strategic.

Unlike chemical explosives, nuclear weapons have had no peacetime uses. In the 1950s, the U.S. government briefly considered using nuclear weapons to blast artificial harbors in the Alaskan coastline but eventually discarded the idea. As of late 2006, nuclear weapons were possessed by a number of nations, including the United States, France, Great Britain, China, India, Israel, Pakistan, the Russian Federation, and North Korea.

The year 2006 was an active one for nuclear proliferation concerns: North Korea exploded its first nuclear device in October 2006. Also during 2006, the United States and European Union jointly asserted that Iran was seeking to develop nuclear technologies that could used in weapons production. Iran denied the claims and asserted that its nuclear programs were designed for the peaceful uses of atomic energy. Using a combination of inspection and remote monitoring (e.g., satellite imagery), The International Atomic Energy Agency (IAEA) in its August 2006 Board summary asserted that that there were no indication of reprocessing activities in Iran but also that Iran has not addressed the long outstanding verification issues or provided the necessary transparency to remove uncertainties associated with some of its activities. Iran has not suspended it enrichment related activities; nor has Iran acted in accordance with the provisions demanded by the international community with regard to transparency.

Since their invention during World War II (19141918), as of 2006 nuclear weapons had been used only twice for non-test purposes. The United States used nuclear weapons against the Japanese cities of Hiroshima and Nagasaki near the end of World War II (19391945). Although U.S. officials at the time asserted that the use of nuclear weapons shortened the war and ultimately saved lives and substantial economic cost for both sides, the decision to use the atomic bomb has long been intensely and emotionally debated. U.S. Army Air Forces at the time of the bombings, Henry H. Arnold, later said that it always appeared to us that, atomic bomb or no atomic bomb, the Japanese were already on the verge of collapse. Moreover, most of the Japanese casualties were civilians.

German physicist Albert Einstein (18791955) did not know it at the time, but when he published his Special Theory of Relativity in 1905 he provided the world with the basic information needed to build nuclear weapons. One aspect of Einstins work (embodied in the famous equation E=mc2) stated that the amount of matter of an object (i.e., its mass) is equivalent to a specific amount of energy. The exact amount of energy in an object equals its mass multiplied by the square of the speed of light. The speed of light is large 186,282 miles per second (300,000 km/sec)so even a small piece of matter contains a vast amount of energy. A baseballsize sample of uranium235, for example, can explode with as much energy as 20,000 tons of TNTand this involves the conversion of only a tiny fraction of the uraniums mass into energy. One pound of explosive material in a fission weapon is approximately 100,000 times as powerful as one pound of TNT. Nuclear fusion weapons, developed in the 1950s, are far more powerful even than this.

As World War II approached, two German chemists, Fritz Strassmann (19021980) and Otto Hahn (18791968), pointed a stream of neutrons at a sample of uranium and succeeded in splitting the nuclei of some of its atoms. This splitting of nuclei is termed nuclear fission. The energy released through nuclear fission was the source of power for the first atomic bomb, which was built in the United States by a large team of scientists lead by U.S. physicist J. Oppenheimer (19041967). This secret research and development program was termed the Manhattan Project.

The first atomic bomb was detonated in Alamogordo, New Mexico, on July 16, 1945. Three weeks later, on August 6, a United States bomber, the Enola Gay, dropped a fourton atomic bomb containing 12 lb (5.4 kg) of uranium235 on the Japanese city of Hiroshima. Seventy thousand people died as a direct result of the blast. Within two months, nearly twice that many were dead from blast injuries and radiation. Three days later, on August 9, a bomb containing several pounds of plutonium was dropped on Nagasaki. Thirty thousand people died in the seconds following the explosion, and more later. The Japanese surrendered the next day, ending World War II.

These first nuclear weapons were atomic bombs or Abombs. They depended on the energy produced by nuclear fission for their destructive power. However, scientists like U.S. physicist Edward Teller (19082003) knew even before the first atomic bomb exploded that the fission weapons could be used to create an even more powerful explosive, now called a thermonuclear device, hydrogen bomb, or Hbomb. This weapon gets it power from the energy released when atoms of the hydrogen isotopes deuterium or tritium are forced together, a process called nuclear fusion. Starting a nuclear fusion reaction is even more complicated than setting off a fission atomic bomb; it requires such heat to initiate it that a fission bomb is used as a detonator to explode the fusion bomb. The United States tested its first hydrogen bomb on November 1, 1952. It exploded with the force of 10.4 megatons (millions of tons of TNT equivalent). Three years later, the Soviet Union exploded a similar device.

For the next 40 years, the United States, with its allies, and the former Soviet Union, with its allies, raced to build more nuclear weapons. Each side produced tens of thousands of nuclear weapons. The end of the Cold War and the breakup of the Soviet Union in the early 1990s led to a significant decrease in the numbers of nuclear weapons in the world; however, the U.S. and the Russian Federation still possess thousands of nuclear weapons.

How nuclear weapons work

Conventional, chemical explosives get their power from the rapid rearrangement of chemical bonds, the links between atoms made by sharing electrons. In chemical explosives, atoms dissociate from other atoms and form new associations; this releases energy, but the atoms themselves do not change. Nuclear weapons are based on an entirely different principle. They derive their explosive power from changes in the structure of the atom itself, specifically, in the core of the atom, its nucleus.

Atomic bombs use the energy released when nuclei of heavy elements split apart or fission. Uranium and plutonium are the two elements that can be used as fuel for this type of weapon. When nuclei of these atoms are struck with rapidly moving neutrons, they are broken into two nearly equal size pieces. They also release more neutrons, which split more nuclei. This is called a chain reaction. If enough atomic nuclei split they will release enough neutrons to ensure that all the nuclei of all the atoms in a sample will be split. Enormous amounts of energy are then released in a fraction of a second. This release of energy is the power behind the atomic bomb.

Uranium and plutonium are termed fissile materials because they can support a fission chain reaction if enough material is concentrated in one place. Too small a sample would not generate enough neutrons to keep the fission process going; for example, a 1lb (.45kg) sample of uranium235, a sample about the size of a pingpong ball, is not large enough to support a chain reaction. The atomic bombs used in World War II proved that 12 or so pounds (about 5.5 kg) of fissile material, larger than a pingpong ball but still small enough to fit into a hand, is enough to maintain a chain reaction. The smallest amount of material that can support a chain reaction is called the critical mass.

The instant enough bomb material is gathered together into a critical mass, a chain reaction begins. (At higher density, less mass is required.) This means that fissile material cannot be assembled in a critical mass until it is meant to explode. Therefore, the sample of uranium or plutonium in an atomic bomb is separated into several pieces, each of which is below critical mass. To set the bomb off, the separated pieces of bomb material are rammed together to create a critical mass. One design for creating a critical mass involves firing a subcritical bullet of fissile material into a subcritical target of fissile material. Together, the bullet and the target create a critical mass that starts a chain reaction leading to a nuclear explosion.

A different design was used to detonate the bomb dropped on Nagasaki. Plutonium was stored in one large but subcritical mass. It was compressed to a critical density by means of surrounding chemical explosives. When the chemical explosive detonated, the blast forced the bomb material into a density that reached criticality. In either type of design, once criticality is reached the explosion follows in a millionth of a second.

In order for nuclear fission to occur, a bomb must use heavy atoms (isotopes) for fuel. Heavy atoms have many nucleonsneutrons and protonsin their nuclei. When these heavy nuclei split apart they release energy (and neutrons, which may cause nearby heavy nuclei to split apart also). Another more powerful type of nuclear weapon uses forms of hydrogen as fuel. Hydrogen has few subatomic particles in its nucleiusually only a proton, but a proton plus a neutron in the isotope deuterium, and a proton plus two neutrons in the isotope tritium. Instead of splitting apart, these light atomic nuclei are forced together in highspeed collisions, a process called nuclear fusion. Energy is released when hydrogen nuclei fuse, forming helium. Fusion only occurs at temperatures of millions of degrees, such as exist in the hearts of stars. (The Sun and other stars generate their energy primarily by fusing hydrogen into helium.) On Earth only an atomic bomb can raise kilograms of material to such a temperature, which is why atomic bombs are used as detonators for hydrogen fusion bombs.

Because hydrogen is lighter than uranium, more hydrogen atoms fit into a sample of the same weight. Thus, even though one fusion reaction releases less energy than one fission reaction, more hydrogen than uranium atoms can be packed into a nuclear weapon and many more fusion reactions can take place in the weapon than fission reactions can take place in a fission bomb. Fusion weapons, therefore, produce bigger explosions than fission weapons of the same physical bulk.

By 1954, a new feature had been added to the hydrogen bomb to create an even more dangerous weapon. Like earlier hydrogen bombs, this weapon was detonated with the explosion of an atomic or fission weapon. This raised temperatures enough to cause the hydrogen atoms in the bomb to fuse and explode like a regular hydrogen bomb. The designers also enclosed this new bomb in a shell of uranium238. Neutrons released from the fusion of hydrogen caused the uranium238 in the surrounding jacket to undergo fission, adding to the power of the blast. This new device was, in effect, a fissionfusionfission bomb.

The power or yield of a nuclear weapon is expressed in terms of how much TNT would be required to equal the weapons blast. Units of kilotons (thousands of tons) and megatons (millions of tons) of TNT are used to describe nuclear blasts.

Effects of nuclear weapons

Nuclear weapons produce two important effects that are also produced by conventional, chemical explosives: they release heat and generate shock waves, pressure fronts of compressed air that smash objects in their paths. The heat released in a nuclear explosion creates a sphere of burning, glowing gas that can range from hundreds of feet to miles in diameter, depending on the power of the bomb. This fireball emits a flash of heat that travels outward from the site of the explosion (ground zero), the area directly under the explosion. This heat can cause second degree burns to bare human flesh miles away from the blast site if the bomb is large enough. (Although this heat can start fires, it seems that much of the fire damage in Hiroshima and Nagasaki following the nuclear explosions resulted from damaged electrical, fuel, gas, and other systems following physical damage caused by the shock or blast wave that accompanied the explosion.)

The explosion of a nuclear weapon creates a shock wave or front of moving air thousands of times more powerful than any produced by any storm, destroying objects in its path. Many nuclear weapons are designed to be detonated high above their targets to take advantage of this shock effect. The more powerful the bomb, the higher in the sky it is usually intended to be detonated. The fission bombs dropped on Japan (Hiroshima, 13.5 kilotons; Nagasaki, 22 kilotons) exploded between 1,500 and 2,000 ft (458610m) above their targets. A bomb with the power of 10 megatons is capable of destroying most houses within a distance of more than 10 miles (16 km) from the blast site.

Unlike conventional explosives, nuclear devices can also release significant amounts of radioactivity and pulses of electromagnetic energy. Radioactivity is the release of fast particles and highenergy photons from unstable atomic nuclei. Besides the greater explosive power of nuclear weapons, radiation is the primary feature that most clearly distinguishes chemical from nuclear explosions. Radiation can kill outright at high doses and cause illnesses, including cancer, at lower doses. The initial burst of radiation during a nuclear explosion is made up of x rays, gamma rays, and neutrons. The energy of this radiation is so high that it can often penetrate buildings. Later, radioactive materials contaminate the explosion site and often enters the atmosphere where it can travel thousands of miles before falling back to earth. This source of radiation is called radioactive fallout. Radioactive fallout can harm living things for years following a nuclear explosion. Fission bombs and fissionfusionfission bombs produce more fallout than hydrogen bombs because the fusion of hydrogen atoms generates less radioactive byproducts than does fission of uranium or plutonium.

Electromagnetic pulses (EMPs) are also produced by nuclear weapons that are exploded at high altitudes, and are caused by the interaction of radiation from the explosion with electrons in the atmosphere and with the Earths magnetic field. EMPs are essentially powerful radio waves that can destroy many electronic circuits.

The effects of fires and destruction following a largescale nuclear war might even change the climate of the planet. In 1983, a group of scientists including U.S. astronomer Carl Sagan (19341996) published the nuclear winter theory, which suggested that particles of smoke and dust produced by fires caused by many nuclear explosions would, for a time, block the Suns rays from reaching the surface of Earth. This, in turn, would reduce temperatures and change wind patterns and oceancurrents. These climatic changes, according to the theory, could destroy crops and lead to the death by famine of many more animals and humans than were killed outright by nuclear explosions. Some scientists have challenged these predictions, but others, including some United States government agencies, support them. On the other hand, there is no controversy about whether a largescale nuclear war could kill hundreds of millions of people and imperil the future of modern civilization, even apart from nuclear winter effects.

Nuclear weapons today

Today nuclear weapons are built in many sizes and shapes not available in the 1940s and 1950s, and are designed for use against many different types of military and civilian targets. Some weapons are less powerful than 1,000 tons of TNT (kilotons), while others have the explosive force of millions of tons of TNT (megatons). Small nuclear shells exist that can be fired from cannons. Nuclear warheads mounted on missiles can be launched from landbased silos, ships, submarines, trains, and large wheeled vehicles. Several warheads can be fitted into one missile and directed to different targets in the same geographic area upon reentry into Earths atmosphere. These multiple independentlytargeted reentry vehicles (MIRVs) can release ten or so individual nuclear warheads far above their targets, making enemy interception more difficult and increasing the deadliness of each individual missile.

In general, nuclear weapons with low yields (that is, in the kiloton, rather than the megaton, range) are termed tactical and are designed to be used in battle situations against specific military targets, such as a concentration of enemy troops or tanks, a naval vessel, or the like. These weapons are termed tactical because the word tactics, in military jargon, refers to the relatively smallscale maneuvers undertaken to win particular battles. Larger nuclear weapons are classed as strategic because the word strategy refers to the largescale maneuvers undertaken to win whole wars. Strategic nuclear weapons are targeted mostly at cities and at other nuclear

KEY TERMS

Atomic bomb An explosive weapon which uses uranium235 or plutonium as fuel. Its tremendous destructive power is produced by energy released from the splitting of atoms or nuclear fission. Also called Abomb, atom bomb, or fission bomb.

Hydrogen bomb An nuclear explosive weapon which uses hydrogen isotopes as fuel and an atom bomb as a detonator. More powerful than an atom bomb, the Hydrogen bomb derives its destructive power from energy released when nuclei of hydrogen are forced together to form helium nuclei in a process called nuclear fusion. Also called Hbomb or Thermonuclear bomb.

Isotopes Two molecules in which the number of atoms and the types of atoms are identical, but their arrangement in space is different, resulting in different chemical and physical properties.

Nuclear fission Splitting the atom. A nuclear reaction in which an atomic nucleus splits into fragments with the release of energy, including radioactivity.

Nuclear fusion A nuclear reaction in which an atomic nucleus combines with another nucleus and releases energy.

Nuclear weapon A bomb or other explosive that derives it explosive force from the release of nuclear energy. PlutoniumA heavy, rare natural element that undergoes fission in a nuclear bomb. It is produced artificially by bombarding uranium238 with neutrons. The addition of one neutron to the nucleus of uranium238 changes it into plutonium239 which is called weapons grade plutonium, the most efficient form for making weapons.

Radioactivity Spontaneous release of subatomic particles or gamma rays by unstable atoms as their nuclei decay.

Radioisotope A type of atom or isotope, such as strontium90, that exhibits radioactivity.

TNT Trinitrotoluene, a high explosive.

Uranium A heavy natural element found in nature. More than 99% of natural uranium is a form called U238. Only U235 readily undergoes fission and it must be purified from the other form.

weapons, and are mostly designed to be dropped by bombers or launched on ballistic missiles; tactical nuclear weapons are delivered by smaller devices over shorter distances. However, one nations tactical warhead may be anothers strategic warhead: Russia, for example, maintains that U.S. tactical warheads in Western Europe are in fact strategic warheads, since they can strike targets inside Russia itself, while Russian tactical warheads in the same arena cannot strike the U.S. heartland.

In the summer of 2002, President George W. Bushs administration sought and received permission from Congress to design a new class of nuclear weapons: mininukes, relatively lowyield tactical nuclear weapons for use against underground bunkers and other small battlefield targets. Advocates of these new weapons point to the uniquely powerful, compact punch that can be delivered by a nuclear weapon; critics argue that even a small nuclear weapon may cause many civilian casualties, and, more important, that actual use of a nuclear weapon of any size would break the taboo on such use that has held since the end of World War II, making the use of larger, more destructive nuclear weapons more likely in future conflicts. Also, nations that currently do not have nuclear weapons might be motivated to obtain them if they saw that the United States or some other power was willing to use them for ordinary warfighting.

Even the ability of nuclear weapons to release radioactivity has been exploited to create different types of weapons. Clean bombs are weapons designed to produce as little radioactive fallout as possible. A hydrogen bomb without a uranium jacket would produce relatively little radioactive contamination, for example. A dirty bomb could just as easily be built, using materials that contribute to radioactive fallout. Such weapons could also be detonated near Earths surface to increase the amount of material that could contribute to radioactive fallout. Neutron bombs originally designed to be used against Soviet forces in areas of Europe with cultural treasures (art museums, etc.) are able to shower battle fields with deadly neutrons that can penetrate buildings and armored vehicles without destroying them. Any people exposed to the neutrons, however, would die. (Neutron bombs also destroy with blast effects, but their deadly radiation zone extends far beyond their blast area)

The United States and Russia signed a Strategic Arms Reduction Treaty in 1993 to eliminate two thirds of their nuclear warheads in ten years. By 1995, nearly 2,500 nuclear warheads had been removed from bombers and missiles in the two countries, according to U.S. government officials. (Elimination, in this context, does not necessarily mean dismantlement; many of the weapons that have been eliminated by treaty have actually been put in storage, not destroyed. As of 2006, approximately 28,000 intact nuclear weapons remained in the possession of Russia and the United States.) Although thousands of nuclear weapons still remain in the hands of many different governments, especially those of the U.S. and the Russian Federation, recent diplomatic trends have at least helped to lower the number of nuclear weapons in the world. This has caused many people to erroneously assume that the danger of nuclear weapons evaporated with the end of the Cold War.

However, the number of nations possessing nuclear weapons continues to increase, and the possibility of nuclear weapons being used against human beings for the first time since World War II may be larger than ever. In May, 1995, more than 170 members of the United Nations agreed to permanently extend the Nuclear Non-Proliferation Treaty, first signed in 1960. Under terms of the treaty, the five major countries with nuclear weaponsthe United States, Britain, France, Russia, and Chinaagreed to commit themselves to eliminating their arsenals as an ultimate goal. The other 165 signatory nations agree not to acquire nuclear weapons. Israel, which many Western intelligence assert possesses some nuclear weapons (but officially denies doing so), did not sign the treaty. Two other nuclear powers, India and Pakistan, also refused to renounce nuclear weapons. India and Pakistan, each of which probably possess several dozen nuclear weapons, have fought a number of border wars in recent decades, and in 2002 came frighteningly close, as many observers thought, to fighting a nuclear war. As mentioned above, North Korea became the worlds most recent confirmed nuclear power in October, 2006.

See also Nuclear power; Nuclear reactor.

Resources

BOOKS

Diehl, Sarah J. and James Clay Moltz. A Handbook of Nuclear Weapons and Nonproliferation. Warwick, UK: Pentagon Press, 2005.

Loeber, Charles R. and Michael Townsend. Building The Bombs: A History Of The Nuclear Weapons Complex. Darby, PA: DIANE Publishing Co., 2004.

Walker, J. Samuel. Prompt and Utter Destruction: President Truman and the Use of Atomic Bombs Against Japan. Chapel Hill, NC: U. of North Carolina Press, 2005.

OTHER

Center for Defense Information. Nuclear Weapons Database. 2003. <http://www.cdi.org/issues/nukeff/database/index.cfm> (accessed November 15, 2006).

K. Lee Lerner

Larry Gilman

Dean Allen Haycock

Nuclear War

views updated May 11 2018

Nuclear War

Some continuities with the past

Differences in technology and doctrine

Deterrence versus denial (or defense)

Changes in prospect

Attitudes toward thermonuclear war

BIBLIOGRAPHY

Nuclear war has but a brief history. In 1945 two low-yield weapons were used against Japanese cities, resulting in the surrender of Japan to the Allies. Nuclear deterrence has had a richer history, but much of this history has little relevance to matters of current importance. This article will therefore focus on the present and prospective consequences of the existence of nuclear weapons.

Some continuities with the past

As context for this discussion, it is important to point out the many factors in the international situation that remain unchanged since 1945. First —and perhaps most important—in spite of international legal institutions that are effective and valuable in a wide range of ordinary situations, crises continue to arise in which each nation becomes the judge of its own cause and of the methods and intensity with which it will seek to advance its interests. Thus, international conflicts have always been prone to “escalation” (a “qualitative” increase in the intensity or level of conflict or violence, perhaps occurring inadvertently, perhaps resulting from a conscious competition in risk taking or even from a deliberate increase in the use of force). I emphasize the pervasiveness of the escalation concept, which implies that in most conflicts there will be force that remains unused. Nuclear weapons have increased the amount of force that tends to remain unused but are not themselves responsible for the possibility that low levels of violence will escalate to medium levels or even very large-scale destruction; that possibility has almost always been present.

Moreover, it is still most unlikely that a thermonuclear war would mean the “end of history,” not because it is technologically impossible to end human history by the use of nuclear explosives but simply because, at this writing (i.e., in 1967), even the superpowers have not procured the kinds of weapons systems that could—in a realistic situation—bring about such a result; nor are they likely to procure such systems. Even if a thermonuclear war were actually fought today by the existing powers, and even if it were fought with the utmost ferocity and lack of control, such a war would be likely to end with the Southern Hemisphere and large portions of Asia largely undamaged. Nor would the survivors in the Northern Hemisphere be likely “to envy the dead.” Recuperation is extremely likely, and, although it is true that according to the best scientific estimates, the postwar environment might be more hostile to human life for many years, objective studies also indicate that this environment would not be so hostile as to preclude, at least in the long run, decent and useful lives for the survivors and their descendants. Indeed, what evidence there is suggests that relatively normal and happy lives would not be impossible even under the harsh conditions that might prevail after a nuclear war, in spite of the personal and social traumas that would have been experienced.

Of course, to say that a nation can survive a thermonuclear war is not to say that all the problems—military, social, political, economic, and medical—that would arise from a thermonuclear war can be handled. Would the social organism fall apart completely—that is, die in some sense— as a result of the tremendous shock it would receive from a large thermonuclear war? Obviously there is no way of being certain. However, insofar as there are historical examples to study (and some of them are close to thermonuclear wars in intensity, e.g., the devastation of Germany and the Soviet Union in World War n), they provide evidence that people can and do rise to the occasion (Kahn 1962).

The age-old role of bargaining in war (e.g., working on the enemy’s mind rather than on his body or resources) may be increased rather than decreased both before and after a nuclear attack. Deterrence is, in effect, a form of bargaining and could occur not only before but also during and after a war (including a war caused by a failure of deterrence)—i.e., even if deterrence fails and war occurs, each side is likely to want to withhold some weapons in order to be able to threaten the enemy and may be anxious to settle the war while the other side still has not used its withheld weapons. The possibility of technological, tactical, or strategic surprise also still exists. Despite current beliefs in stability, “Pearl Harbors” and “Munichs” remain possible (i.e., deterrence may not necessarily favor the defensive or status quo side); thus, those who take risks may find that their gamble has not failed. Finally, even though countries will be less likely to risk war, in some ways surprises and “unexpected” victories are more likely in a nuclear war era (i.e., war could occur as a result of correct calculations as well as by miscalculation).

While mutual destruction and Pyrrhic victories can occur, such possibilities have been with us at least as long as recorded history, and the fact that they could occur days or even hours after the start of hostilities, rather than after weeks, months, or years, does not necessarily change the basic issues and situations that have arisen in the past. Thus, history may repeat itself in the thermonuclear era, which could experience successful wars and aggressions as well as mutually destructive ones.

Differences in technology and doctrine

The most important way in which the thermonuclear era differs from previous eras is that the changes in the available technology have, by themselves, been large enough to change the strategic situation and the character of international relations. This has resulted both in posing very different strategic issues and in changing the answers to old issues. Thus, in this period U.S. strategic weaponry has changed (or is changing) from B-50 bombers to B-36 bombers, to B-47s, and finally to B-52s; and from Atlas and Titan intercontinental ballistic missiles to early-model Minuteman and Polaris, and now to Minuteman-2s and Polaris A-3s. These changes have produced enormous differences in the effectiveness of the forces. Even the change from Minuteman-1 to Minuteman-2 can by itself increase the effectiveness of a force by a large factor (doubling—or perhaps redoubling several times—the size of a nation’s missile forces), to the extent of altering the outcome of a war.

However, since all these changes have occurred without being tested in battle, one can question whether their full significance has even been absorbed intellectually, much less in plans and programs.

Indeed, it is clear that in many respects the two superpowers, and certainly the other nations, have failed to understand these rapidly changing interactions between technology and doctrine. We can (with some violence to the subtleties, and viewing the subject from the viewpoint of the U.S. government) consider the two postwar decades as four five-year eras characterized as follows:

1946-1950: Early fission and bomber era; entry of Soviet Union into nuclear club.

1951-1955: Early thermonuclear weapons; maximum “apparent” U.S. superiority; simple massiveretaliation theories; entry of United Kingdom into nuclear club.

1956-1960: Early missile-middle thermonuclear weapons era; intense debates about deterrence and tactical nuclear war; continuation of (and probably increased but less apparent) U.S. superiority.

1961-1965: Middle missile era; two-day deterrence; controlled-war doctrine; entry of France and China into nuclear club.

In the early nuclear bomber era, the strategic balance was one-sided, for the Soviets did not test their first weapon until February 1949. However, the United States had not produced very many nuclear bombs, and it is doubtful that the U.S. strategic forces could have had anything like the decisive impact on a war that was popularly supposed—indeed, these forces might have done less damage to a Soviet war effort than the Germans did in World War n by invading and occupying Soviet territory.

By the early 1950s the United States had substantially increased its forces, but the Russians had scarcely begun to procure intercontinental bombers. They had a rather large force of medium bombers of the TU-4 (similar to the B-29) and Badger (like the B-47) types, but it now seems clear that both of these aircraft were designed and procured for European rather than intercontinental missions—although at the time no one in the United States or Europe seemed to realize this. Although thermonuclear weapons had been tested, the military stockpiles consisted almost wholly of kiloton bombs.

Despite slow means of delivery (bombers) and relatively low-yield warheads, both U.S. and Soviet forces were almost incredibly vulnerable to surprise attack. At the beginning of the period all U.S. strategic forces were located at a dozen bases. Hours, perhaps days, would have been required to evacuate them and days, perhaps weeks, of warning would have been necessary for them to be able to mount effective combat operations. Nuclear weapons were stored in a relatively vulnerable configuration (at first in one building and then in two). In the early part of this period, almost no one seems to have understood the subtleties of the problem of vulnerability. Active air defense was deployed to protect cities and nuclear research centers like Oak Ridge, Tennessee, and Hanford, Washington. Strategic Air Command bases were left unguarded, on the theory that no one would waste nuclear weapons on military bases. By the middle and end of the period, senior officers in the U.S. Air Force understood vulnerability somewhat but did not believe (it now seems correctly) that the Soviets had much actual operational strategic capability for a surprise attack.

By the late 1950s, third- and fourth-generation nuclear weapons had been procured by the United States and a large spectrum of such weapons was available to the U.S. forces, from small (“suitcase”) to multimegaton bombs. U.S. military planners and decision makers began to think of thermonuclear weapons as relatively inexpensive, but the Soviets still did not. However, the problems of vulnerability were still inadequately appreciated; in fact, by the end of the period there was much discussion of the existence of a “missile gap.” While the U.S. government conceded the existence of such a gap—and in fact was responsible for disseminating the estimates which gave it plausibility—it simultaneously argued there was no “deterrence gap,” since the five hundred missiles which U.S. intelligence attributed to the Soviet Union would have been unable to do as much damage as the two thousand bombers that the United States possessed. Congressional testimony and other documents disclose that almost none of the top civilian officials and relatively few scholars and journalists understood that these five hundred Soviet missiles, if they existed, could probably have destroyed the two thousand American bombers on the ground in a surprise attack (but see Wohlstetter 1959; Kahn 1960).

By the early 1960s the United States, at least, was well into the missile era, and almost everybody interested in such problems understood the distinction between “first-strike” and “second-strike” tactics, forces, and postures. However, according to the 1963 testimony of Defense Secretary Robert McNamara, the Soviets had not yet hardened and dispersed their missile forces, although it was expected that they would do so by the late 1960s. In the early 1960s some of the doctrinal lags of the late 1950s were revealed. For example, it was disclosed that the more important half of the U.S. Semi-automatic Ground Environment Air Defense System, the part designed to control the air battle in defense of strategic centers on the American east coast, west coast, and Canadian border, was located on Strategic Air Command (SAC) bases and thus was almost certain to be destroyed or disabled in any war in which Soviet missiles were successfully launched at these SAC bases. Similar mistakes in both installations and weapons systems occurred elsewhere. As a result, there was tremendous concern about vulnerability and emphasis on such problems as “reciprocal fear of surprise attack” (expressed in terms of “gun duel” models of a strategic confrontation in which the side which gets off a “shot” first may escape all retaliation).

By the end of this period the United States had begun to digest not only the preattack implications of two-way deterrence but also the possibility of intrawar or postattack deterrence, and therefore the need for restraint in the threat and use of force even after hostilities have begun. This resulted in the so-called controlled-response doctrine (Kahn 1960, pp. 171-175) and such policy statements as the one by Secretary McNamara that “principal military objectives in the event of a nuclear war stemming from a major attack on the Alliance, should be the destruction of the enemy’s military forces, not of his civilian population”(New York Times, June 17, 1962, pp. 26-27) and the statement of President Johnson in his defense message to the 89th Congress that “our military forces must be so organized and directed that they can be used in a measured, controlled, and deliberate way as a versatile instrument to support our foreign policy” (1965, p. 825).

Despite these statements, however, it becomes quite clear from other statements that both Secretary McNamara and President Johnson remained doubtful about the feasibility, likelihood, or even possibility of a controlled response in a major war. Moreover, there is little indication that the United States has thoroughly organized its forces (not to speak of the NATO forces) around these concepts.

The change from kiloton to megaton weapons in the early 1950s was, in some ways, as significant as the change from high-explosive to kiloton weapons in the mid-1940s. Until megaton weapons became available, it was unlikely that the U.S. Strategic Air Command could have done as much physical damage to the Soviet Union’s war effort as was actually accomplished by the German army in World War n. In the mid-1960s, in the multimegaton era, potential destruction from a thermonuclear war is almost incredibly large. The duration of an “all-out” war against city targets is now likely to be closer to a minimum of 30 minutes than to a maximum of 30 days. Analysts no longer ask “What is destroyed?” but “What is left?” What has also become apparent, loose rhetoric to the contrary, is that thermonuclear wars may come in many sizes and shapes. The analysts must, consequently, examine and estimate a large number of effects in a wide range of situations.

First, prewar tactics and strategy, which may make an enormous difference to the outcome, must be studied. Then the immediate effects of blast and “prompt” gamma and thermal radiation must be examined, as well as the subsequent effects of primary and secondary fires and of fallout radiation during the first week or two. If the war has involved widespread targeting of civilian areas, there are problems of human survival, of radioactivity, and of damage to the physical environment, and problems of economic and social disorganization. These require consideration of immediate repair and “patch-up,” and economic and social reorganization and recuperation (problems which so far have hardly been dealt with, despite many millions of dollars spent on research in these areas). Even in a controlled thermonuclear war the likely rate of economic and social recuperation (assuming the recuperation efforts themselves have been successfully launched) may be very different from those which would follow even a very destructive conventional war. There are also long-term environmental problems, including the medical aftereffects of radiation on exposed survivors. Finally, account must be taken of the genetic effects of radiation and long-range changes in the physical environment resulting from widespread radiation damage to plants and animals, large-scale fires, floods, possibly genetic changes in flora and fauna, and even weather changes.

Many of these questions are unprecedented; all are complex; and almost all tend to arouse “openended” fears. But in some ways the appearance of total uncertainty on these issues is misleading, since the results of thermonuclear war, particularly the destruction of enemy weapons and, to some degree, cities, as well as many physical aftereffects, are easier to calculate than are the results of the clash of two land armies fighting a conventional war. However, the lack of experience in such wars, and the absolute necessity for making and relying on decisions and analyses made in advance, means that, from a decision-making point of view, the situation is much more complex and uncertain. Also, while one can often make relatively precise estimates about the probability that something will be destroyed by some specific mechanism, it is often impossible to make precise statements that something has survived all the ways in which it could be destroyed.

Deterrence versus denial (or defense)

One result of the potential increase in destructiveness combined with confusing complexities and uncertainties has been a major change in attitudes toward war. Such phrases as “absolute weapon” (doomsday machine), “balance of terror,” “live together or die together,” and “war is unthinkable” (or impossible or obsolete) illustrate widely held and widely expressed attitudes and beliefs. Not only is such total destruction hard to visualize, but it leads to the feeling that the deterrence of war, or even of the threat of war, is a simple and logical consequence of the existence of nuclear weapons.

The total destruction or “mutual homicide” interpretation of thermonuclear war has other comforting aspects. If it be granted that each side can utterly and reliably destroy the other, then expensive preparations to reduce casualties, lessen damage, and facilitate postwar recuperation are useless. Can we not spare ourselves the financial burden of such preparations? This logic has sometimes been carried further, to argue that if modern weapons are so enormously destructive, then only a few are needed to deter the enemy. War can be deterred with much smaller forces than in the past; and we certainly don’t need larger ones. Most influential of all of these arguments is that if destruction is always total and automatic, we need not spend time and energy worrying about details, comparison of risk, etc.

All of the above has led to a well-articulated and explicit dependence on deterrence—on dissuasion through terror—and the belief that the major role of a nuclear force is that of a deterrent and bargaining tool. As President John F. Kennedy said on March 28, 1961, in his special message to the Congress on the defense budget:

The primary purpose of our arms is peace, not war—to make certain that they will never have to be used—to deter all wars, general or limited, nuclear or conventional, large or small—to convince all potential aggressors that any attack would be futile—to provide backing for diplomatic settlement of disputes—to insure the adequacy of our bargaining power for an end to the arms race. (New York Times, March 29, 1961 pp. 16, 17)

Whether based on objective capabilities or “resolve,” the concept and use of deterrence are not new. However, today there is almost total emphasis on mutual terror and destruction—that is, the countervailing power does not emphasize its ability to negate the acts of the aggressive or active power, or to destroy or block his forces militarily, as in the past, but emphasizes instead its ability to harm the population, resources, or property of the opposing power. An explicit distinction is made today between deterrence and what is sometimes called denial (or defense), which is the physical prevention or alleviation of an action (as opposed to a “psychological” prevention, based mainly upon the threat of pain or destruction of other values). The relationship between deterrence and denial—the various ways in which they can reinforce or conflict with each other—has become the subject of much discussion (e.g., Kaufmann 1956; Kahn I960; 1962; Schelling I960; 1966).

Deterrence is, of course, a complicated relationship, and in trying to analyze deterrence situations, one may elaborate on Raymond Aron’s well-known questions by asking:... Who deters, influences, coerces or blocks whom from what actions (alternatives), by what threats and counteractions in what situations and contexts, in the face of what counterthreats and counter-counteractions?... and why does he do it? .. .

One can group the italicized variables into three groups: (1) political (who, whom, and why); (2) scenario (alternatives, situations, and context); and (3) military (actions, threats [or counteractions], and counterthreats [or counter-counteractions]). While most discussions of thermonuclear war generally emphasize only one or another of the above sets of variables, all three sets must be considered in an integrated way. (Only two works seem to have attempted this: Kissinger 1957; Strausz-Hupe et al. 1959.) Failure to do so constitutes a major source of misunderstandings about deterrence.

It is important to distinguish between what is sometimes called “passive” and “active” deterrence. “Passive” deterrence refers to a situation in which a nation has been so provoked (perhaps by a direct attack on its population) that the response is almost automatic. “Active” deterrence applies to a situation in which the “proper” response to provocation is clearly further escalation (perhaps one’s ally has been struck). In this situation it takes an act of will to respond. Rather than being an automatic act of revenge, the response may start a sequence leading to one’s own destruction.

It is also a mistake to treat deterrence as an either/or situation, instead of considering degrees of deterrence. In most situations one could usefully distinguish at least six levels of deterrence between the United States and the Soviet Union (Kahn 1965, pp. 277-280):

(1) Minimum (or low-level): (a) deterrence by uncertainty (including the use of dire but unbelievable threats as a declaratory policy but preemptive or preventive accommodation tactics as an action policy); (b) deterrence by threshold; (c) deterrence by taboo.

(2) Workable: reliable ability to threaten, say, 1-10 million casualties or unreliable ability to threaten, say, 10–50 million casualties.

(3) Adequate: reliable ability to threaten, say, 10-50 million casualties or unreliable ability to threaten, say, 50-100 million casualties.

(4) Reliable: reliable ability to threaten, say, 50-100 million casualties.

(5) Approaching absolute: reliable ability to threaten, say, 100-200 million casualties.

(6) Near absolute (or stark): reliable ability to threaten overkill by a factor greater than 2, so that no miscalculation or wishful thinking could confuse the potential attacker.

Although the above scale leaves out most of the subtleties associated with, for example, credibility issues, it correctly implies that there is a broad range of circumstances in which even a minimum deterrent might work and that there are circumstances in which the most stark deterrent might fail. The scale also suggests that the question “How much deterrence does a nation need?” cannot be answered simply by “As much as possible.” Rather, a range of scenarios, asking the whowhom-why kind of question not discussed here, must be examined, and, most important of all, the degree of assurance necessary for various situations must be weighed against the various costs of going higher on the scale of deterrence.

Deterrence clearly involves dissuasion by use of threat of varying degrees and kinds of terror, which in turn assumes that the deterree (and even the deterrer) possesses some degree of rationality— but usually very little: about as much as is demonstrated by a child who has learned not to burn himself or climb out of windows. Nevertheless, the emphasis on terror and rationality raises much apprehension, particularly among members of the peace groups but also among decision makers and analysts generally. This is not only because subtle and misleading situations can arise but also because many people feel, often reasoning from the psychology of the neurotic and psychologically disturbed, that threats and terror may attract rather than repel and that rationality is thus a weak reed on which to rest our hopes for preventing thermonuclear war. One can, however, fully share these apprehensions and still feel that deterrence is, for the time being, the safest and most moral alternative available. But while apprehensions are often exaggerated, even the most optimistic may still wish some protection against the failure of deterrence and may work to improve and eventually to reform the deterrent system. [SeeDeterrence.]

Changes in prospect

By the late 1960s at least five or six countries should have reached India’s position in 1965, i.e., that of having completed almost all research necessary to assemble and test a nuclear device but having stopped short of developing a “working” model (a matter of perhaps less than a year and less than a $1 million cost) because of governmental reluctance to authorize this final step. This situation may become typical, and it would not surprise most scholars if there were no new entries into the “nuclear club” during the next decade. As of 1966, there have been no public entries into the nuclear race for more than ten years: China and France made and announced their initial decisions to become nuclear powers before this period.

In terms of delivery systems, the late 1960s will be the early missile era, in that most nations will not have achieved the kind of capabilities that the United States will possess. The United States, in turn, will probably have entered the mature missile era, since at least some U.S. missiles will be cheap, reliable, and relatively small in size; will possess great range, good accuracy, and reasonable payload capability; and will be capable of complicated or sophisticated tactical performance.

The cost of a simple Minuteman missile system is likely to be something between $1 and $2 million a year per missile to buy and maintain, as long as there is reasonable access to production and operating techniques comparable to those of the United States. These costs are likely to apply to missile systems with at least hundreds but not necessarily thousands of missiles. Thus, on the basis of U.S. costs, any nation will be able to have, say, a 500missile Minuteman-type strategic force for a budget of between $500 million and $1 billion a year.

There are many who now believe that the kind of revolutionary technological changes that have occurred since 1945 will, for practical purposes, have come to a stop. This belief is usually based on various versions of overkill theories and the conviction that nuclear “stalemates” are not susceptible to breakthroughs in technology. The argument runs as follows:

From 1965 on, or soon afterward, both the United States and the Soviet Union will be able to inflict unacceptable damage on each other under all circumstances, using more or less existing, or moderately improved, systems. Thus, if one nation or the other provides a new way to damage or destroy its potential opponent’s civilian society, the deterrent equation is not affected. For this reason the many possible technological improvements in offensive power (against civilians) will not affect international relations. The same will probably be true in the unlikely event that there are major improvements in defense (or counterforce) capability. The current stalemate is at a high enough level so that each nation will still be able to inflict unacceptable damage in retaliation, even if its capability is cut by a large factor. This will be true practically, even if it is not true theoretically, because if one side does invent and procure some breakthrough, it cannot rely on the system’s working properly (perfectly?) the first time it is put to the test and will therefore be “deterred by uncertainty,” and deterred almost as effectively as if the deterrence rested on objective capabilities and calculations. Therefore, at least in the area of central war (i.e., nuclear war involving the homelands of the major powers), there is little or no interest in examining technology with the same intensity that was necessary in the last two decades.

Even though it ignores many important issues, this argument is persuasive and may prove to be practically correct as far as the politics of U.S. Soviet confrontations are concerned—although both countries will still work at improving their forces, and such improvements might make a dramatic difference in the outcome if deterrence ever failed. In any case, as Figure 1 shows, budgets for central war forces have been declining over the years (in the late 1950s the total budget was probably close to $15 billion). The data suggest either that thermonuclear war is getting much less expensive or that the United States is becoming satisfied with much less capability (actually, both effects are occurring).

The argument leaves open the question of a confrontation between one of these central superpowers and an nth country or a confrontation between nth countries. In these situations many of the points made above lose much of their force. Although such lesser confrontations do not seem to raise the question of the future of the entire world as immediately and intensely as U.S.-Soviet ones, the most significant technological arms race will probably go on “outside” the direct confrontation of the two superpowers (unless, of course, the superpowers go in for elaborate arms control or active and passive defense, in which case technology will again make a great difference). There will in any case be a widespread distribution of longrange missile technology using both solid and liquid fuels. Missile technology will proliferate more rapidly than weapons technology because of lesser secrecy, due, in part, to the much greater application of missile technology to normal peacetime research and engineering.

The strategic situation will probably be very much affected by whether or not the United States and/or the Soviet Union allocate sufficient resources to active and passive defenses to keep ahead of the later entries into the nuclear club. If either spends, say, $5-$ 10 billion a year in this area, they will probably preserve an important, perhaps overwhelming, strategic asymmetry. If they do not do this, and the likelihood is that they

* Includes procurement, maintenance, and operation of strategic offense and active and passive defense of the United States, but no R & D programs.

will not, the situation will be closer to that illustrated by an example drawn from the American West, where the six-shooter was the great equalizer, and there will be a lesser tendency to differentiate between superpowers and great powers.

If there is widespread proliferation, the situation will doubtless be quite dangerous; but despite the obvious dangers, it is not likely that a war between two small or medium powers, or even large powers, would trigger an orgy of destruction in which every power joins. In a world of nuclear proliferation the concept of controlled response will be firmly fixed in everybody’s mind, and it is inconceivable that many, if indeed any, nations will want to enlarge a nuclear war in which they are not directly and vitally involved—and perhaps not even then. It may even be that a nuclear war may occur between two smaller powers in which both sides are wiped out, or even in which one side “wins” but at a cost which results in a sobering, not to say chilling, effect on all. Most dangerous of all, one power may win clearly and easily, and succeed in keeping its winnings. This might be a great spur both to further proliferation and to risk taking by some of the more aggressive nuclear powers.

By the late 1970s, in short, the technological and economic possibilities for proliferation will be such that unless there are extreme implicit or explicit constraints against it, a widespread or even explosive proliferation can be expected. By the 1980s we will be either in the postnuclear era or in what may be thought of as the mature nuclear era. That is, we will be living either in a world in which large numbers of powers have nuclear weapons or in a world in which nuclear weapons have been effectively controlled.

Attitudes toward thermonuclear war

Parallel to the acceleration of the advances in modern technology in the twentieth century, there has been a growing feeling that technological leaps forward (rather than morality) have made war unacceptable. However, another reason why war is “unacceptable” should also be noted: the existence of a “peace of satisfaction” (Aron 1962). To illustrate this phrase, consider Latin America today, where for the last two decades there has been almost no major threat of war. The explanation obviously has nothing to do with nuclear weapons. Yet if these nations possessed such weapons, almost everyone would ascribe the relative peace of the period to the existence and effectiveness of nuclear deterrence. Similarly, most of the large nations in the world today are deterred—but even more, they are defensive, prudent, and in large part supporters of the status quo. Nevertheless, nuclear weapons do pose the possibility of mutual destruction and thus raise the question of war’s practicality or usefulness as a social institution in an entirely different way than does “conventional” war. The answers to this question reflect at least three common attitudes toward nuclear weapons:

(1) An unqualified rejection, which also includes a rejection of deterrence —i.e., it is argued that since deterrence can fail, we must move to a situation in which the weapons themselves no longer exist. This condition would doubtless be desirable, but it is difficult to see how it is likely to be achieved without much more peaceful evolution—or crises or wars—than seems likely in the immediate future. It seems unlikely that the threat of nuclear war will be “settled” through peaceful evolution and international cooperation in the next decade or so, although these may help facilitate the “eventual settlement.”

(2) A more common attitude of qualified rejection of nuclear weapons as useful for any purpose except to deter their use by others. All the various kinds of finite deterrence strategies and basic deterrence strategies express this attitude, and little or no preparation is advocated for alleviating the consequences if deterrence fails.

(3) A qualified acceptance of nuclear weapons or resignation to their continued existence and possible use. Acknowledging all of the points made earlier on ways in which the world situation has not changed, some of these “acceptors” believe in a self-defeating prophecy—i.e., that deterrence will be much more likely to fail if too many people believe that “war is unthinkable.” Some of this group also argue that even during a war there might be deterrence of many actions and thus some degree of control. Few of this group would argue that such intrawar deterrence can be relied on. All believe, however, that the likelihood of maintaining control, even after nuclear weapons have been used, is so high that attempts should be made to exploit it. Finally, there are those who wish also to support a bargaining position, or to hedge, against a loss of control, by having some active and passive defense and some offensive counterforce ability.

Whatever the attitude about the feasibility of various solutions, hardly anyone considers the arms race or large-scale war useful in international relations. President Kennedy expressed, in two aphorisms, what seems to be the consensus of mankind. The first was in his UN address of September 25, 1961: “The weapons of war must be abolished before they abolish us”; the second, in his budget message of January 18, 1962, was that we must “retain for ourselves a choice other than a nuclear holocaust or retreat.” The first statement counsels us to change the current international system; the second, to maneuver successfully within it. The dilemma of the nuclear age is usually posed by arguing that these two imperatives are both valid and mutually exclusive. And indeed, although it may prove impossible to change the current system until after it has failed conspicuously and perhaps disastrously, the second imperative could also help to deal with the first one.

In any case, we seem, for the time being, to be succeeding in avoiding both disastrous appeasement and withdrawal or escalation to catastrophic levels. To hope that this success can be perpetuated is to hope for more than just the accurate and detailed awareness of the need to maintain military and political strength while limiting the new potential for unprecedentedly rapid, long-range, and widespread destructiveness. These efforts, of course, are more likely to avert eruption of a cataclysmic war or appeasement only if there is understanding of all the dangers and options. In addition, increased awareness of all the issues may motivate sufficient international reform and simultaneously constrain all national decision makers to exhibit a consistently high degree of caution, reasonableness, and restraint in their dealings with one another, whether in peace or in limited war. The issue is, of course, still open.

Herman Kahn

[See alsoDeterrence; Disarmament; Limited War; Technology, article onTechnology and International Relations. Guides to other relevant material may be found underInternational Relations; War.]

BIBLIOGRAPHY

Aron, Raymond (1958) 1959 On War. Garden City, N.Y.: Doubleday. → First published as De la guerre.

Aron, Raymond (1962) 1967 Peace and War: A Theory of International Relations. Garden City, N.Y.: Doubleday. → First published as Paix et guerre entre les nations.

Aron, Raymond (1963) 1965 The Great Debate: Theories of Nuclear Strategy. Garden City, N.Y.: Doubleday. First published as Le grand debat: Initiation a la strategie atomique.

Beaton, Leonard; and Maddox, John R. 1962 The Spread of Nuclear Weapons. New York: Praeger.

Beaufre, Andre 1965 An Introduction to Strategy. New York: Praeger.

Berkowitz, Morton; and Bock, P. G. (editors) 1965 American National Security: A Reader in Theory and Policy. New York: Free Press.

Brennan, Donald G. (editor) 1961 Arms Control, Disarmament, and National Security. New York: Braziller.

Brodie, Bernard 1959 Strategy in the Missile Age. Princeton Univ. Press.

Brodie, Bernard 1966 Escalation and the Nuclear Option. Princeton Univ. Press.

Brown, Neville 1965 Nuclear War: The Impending Strategic Deadlock. New York: Praeger.

Bull, Hedley (1961) 1965 The Control of the Arms Race: Disarmament and Arms Control in the Missile Age. 2d ed. New York: Praeger.

Dinerstein, Herbert S. 1959 War and the Soviet Union: Nuclear Weapons and the Revolution in Soviet Military and Political Thinking. New York: Praeger.

Falk, Richard; and Mendlovitz, Saul H. (editors) 1966 The Strategy of World Order. 4 vols. New York: World Law Fund.

Gallois, Pierre (1960) 1961 The Balance of Terror: Strategy for the Nuclear Age. Boston: Houghton Mifflin. → First published as Strategie de Iàge nucliaire.

Gilpin, Robert (1962) 1965 American Scientists and Nuclear Weapons Policy. Princeton Univ. Press.

Halperin, Morton H. 1963 Limited War in the Nuclear Age. New York: Wiley.

Hitch, Charles J.; and Mckean, Roland N. 1960 The Economics of Defense in the Nuclear Age. Cambridge, Mass.: Harvard Univ. Press.

Hoffmann, Stanley 1965 The State of War: Essays on the Theory and Practice of International Politics. New York: Praeger.

Ikle, Fred C. 1961 After “Detection”—What? Foreign Affairs 39:208-220.

Johnson, Lyndon B. 1965 National Defense Message From the President of the United States. U.S. Congress, Congressional Record III, part 1:823-827.

Kahn, Herman (1960) 1961 On Thermonuclear War. 2d ed. Princeton Univ. Press.

Kahn, Herman 1962 Thinking About the Unthinkable. New York: Horizon.

Kahn, Herman 1965 On Escalation: Metaphors and Scenarios. New York: Praeger.

Kaufmann, William W. (editor) 1956 Military Policy and National Security. Princeton Univ. Press.

Kissinger, Henry A. 1957 Nuclear Weapons and Foreign Policy. New York: Harper.

Kissinger, Henry A. (editor) 1965 Problems of National Strategy: A Book of Readings. New York: Praeger.

Knorr, Klaus E.; and Read, Thornton (editors) 1962 Limited Strategic War. Published for the Center of International Studies, Princeton University. New York: Praeger.

Lyons, Gene M.; and Morton, Lewis 1965 Schools for Strategy: Education and Research in National Security Affairs. New York: Praeger.

Millis, Walter; Mansfield, Harvey C ; and Stein, Harold 1958 Arms and the State: Civil-Military Elements in National Policy. New York: Twentieth Century Fund.

Ramsey, Paul 1961 War and the Christian Conscience: How Shall Modern War Be Conducted Justly? Durham, N.C.: Duke Univ. Press.

Rosecrance, Richard N. (editor) 1964 The Dispersion of Nuclear Weapons: Strategy and Politics. New York: Columbia Univ. Press.

Schelling, Thomas C. 1960 The Strategy of Conflict. Cambridge, Mass.: Harvard Univ. Press.

Schelling, Thomas C. 1966 Arms and Influence. New Haven: Yale Univ. Press.

Schmidt, Helmut (1961)1962 Defense or Retaliation: A German View. New York: Praeger. → First published in German.

Snyder, Glenn H. 1961 Deterrence and Defense: Toward a Theory of National Security. Princeton Univ. Press.

Sokolovskii, Vasilii D. (editor) (1962) 1963 Military Strategy: Soviet Doctrine and Concepts. Introduction by Raymond L. Garthoff. New York: Praeger. → First published in Russian.

Strausz-Hupe, Robert et al. 1959 Protracted Conflict. New York: Harper. → A paperback edition was published in 1963.

Tucker, Robert W. 1960 The Just War: A Study in Contemporary American Doctrine. Baltimore: Johns Hopkins Press.

U.S. Air Force Academy, Department of Political Science 1965 American Defense Policy. Prepared by Wesley W. Posvar and others. Baltimore: Johns Hopkins Press.

U.S. Armed Forces Special Weapons Project (1950) 1962 The Effects of Nuclear Weapons. Edited by Samuel Glasston. Washington: U.S. Atomic Energy Commission. → First published as The Effects of Atomic Weapons.

Wohlstetter, Albert 1959 The Delicate Balance of Terror. Foreign Affairs 37:211-234.

Wohlstetter, Albert (1961) 1965 Nuclear Sharing: NATO and the N + 1 Country. Pages 186-212 in Henry A. Kissinger (editor), Problems of National Strategy: A Book of Readings. New York: Praeger. → First published in Volume 39 of Foreign Affairs.

Nuclear Weapons

views updated May 23 2018

Nuclear Weapons

Genocide and crimes against humanity can be carried out with machetes. They can be carried out with nuclear weapons. It appears, however, that in the current state of international law, using a nuclear weapon on people may not, in itself, be genocide, a crime against humanity, or otherwise absolutely forbidden.

The Nuclear Age arrived in the desert near Los Alamos, New Mexico, on July 16, 1945, with the first nuclear test detonation. That same year, the bombs were dropped from U.S. planes on the Japanese cities of Hiroshima, on August 6, and Nagasaki, three days later. Hiroshima and Nagasaki have been the only uses of nuclear weapons in armed conflict.

Subsequently, the Soviet Union, the United Kingdom, France, and China became avowed members of the Nuclear Club. The United States and the Soviet Union tested hydrogen devices of ever more awesome power. Israel is widely believed to have nuclear weapons. South Africa probably had the capability but fors-wore development after the demise of apartheid. India and Pakistan tested devices in 1998. North Korea apparently has the capability, and Iran, Iraq, Libya, and Brazil have been suspected of developing it. Iraq's nuclear potential was a significant factor in the efforts by the United Nations and the International Atomic Energy Agency (IAEA) to control that country's development of weapons of mass destruction following the Gulf War of 1991. (The IAEA is an intergovernmental organization associated with the United Nations that is devoted to encouraging peaceful uses of nuclear energy.) Nuclear potential figured prominently in the rationale articulated by the United States for its pre-emptive invasion of Iraq in 2003.

Nuclear weapons are explosive devices whose energy comes from fusion or fission of the atom. Their explosion releases vast amounts of heat and energy as well as immediate and long-term radiation. Radiation, unique to nuclear weapons, can cause nearly immediate death and long-term sickness, as well as genetic defects and illness in future generations. Nuclear weapons can have dramatically greater explosive effect than conventional weapons. The bomb dropped on Hiroshima from the airplane named Enola Gay was the explosive equivalent of approximately three thousand B29 bombers carrying conventional bombs. The "Bravo" hydrogen test at Bikini Atoll in 1954 had one thousand times the power of the Hiroshima blast.

The Case of Hiroshima and Nagasaki

What might international law say of such forces? The first legal assessment came as a protest from the Japanese Imperial Government through the Government of Switzerland, four days after the bombing of Hiroshima. Referring to Articles 22 and 23 (e) of the Regulations respecting the Laws and Customs of War on Land annexed to the Hague Convention of 1907, the Japanese government emphasized the inability of a nuclear bomb to distinguish between combatants and belligerents, and the cruel nature of its effects, which it compared to poison and other inhumane methods of warfare. Article 22 of the Hague Regulations provides: "The right of belligerents to adopt means of injuring the enemy is not unlimited." Article 23 (e) provides that ". . . it is especially forbidden . . . (e) To employ arms, projectiles, or material calculated to cause unnecessary suffering." The Japanese protest decried "a new offence against the civilization of mankind." Its adversary, however, emphasized how the use of the bomb had quickly brought the war to finality, with millions of lives saved by avoiding a sea and land assault on the Japanese mainland.

The Japanese protest appears as Exhibit III in Shimoda v. State, a case brought in the Tokyo District Court in 1963. The plaintiffs sought damages for injuries suffered in Hiroshima and Nagasaki. The plaintiffs argued the illegality of the use of nuclear weapons, founded on an expanded version of the 1945 protest. Damages were claimed from the Japanese government on the theory that it had, in the Peace Treaty, waived the rights of victims to obtain redress from the United States without supplying an alternative source of compensation. The court agreed that the bombings were illegal, but held there was no right to press a claim for damages against the Japanese government.

The concepts of genocide and crimes against humanity were not yet in wide usage when the Japanese government made its August 1945 protest. If the events had occurred a little later, after the concepts gained currency, the government might have added references to those concepts in its protest. Given the international conflict with the United States, however, it was natural to rely on the law of the Hague.

The general principles of the laws of armed conflict have been a major recurring theme in the efforts to rein in nuclear weaponry through international law. This strategy emphasizes banning the use (but not necessarily possession) of such weapons. Other means have included: the quest for partial or total nuclear disarmament (including efforts at non-proliferation and strategic arms limitation); attempts by treaty, resolutions in international organizations, and litigation to stop the testing of such devices; limitations on the development of delivery systems (and defenses thereto); and the creation of Nuclear Free Zones, such as Antarctica, the moon, the South Pacific, and Latin America.

Australia/New Zealand Law Suits

New Zealand incurred the wrath of its traditional allies in the 1980s by instituting a total ban from its ports of nuclear-armed and nuclear-powered vessels. In 1973 Australia and New Zealand endeavored to obtain a ruling from the International Court of Justice on the legality of French nuclear tests in the Pacific. Their arguments relied primarily on environmental law and the law of the sea. A majority of the court in effect held the case moot, without reaching a finding on the merits. France had, until the time of the proceedings, been testing in the atmosphere. It now announced that its future tests would be underground. The court held that this announcement was legally binding on the government, which meant that the object sought by Australia and New Zealand had been achieved.

The court, in vague language, left open the possibility of revisiting the case "if the basis of this Judgment were to be effected." New Zealand believed that its case dealt not only with tests in the atmosphere, but also tests that resulted in the entry of radioactive material into the marine environment, even if the testing took place below the ground. Receiving indications that radioactive material was escaping from underground, New Zealand tried to resurrect its case in 1995. A majority of the court refused to reopen the case, taking a narrow view of the earlier proceedings and insisting that, like Australia's somewhat differently worded case, only atmospheric testing had been at issue.

International Court of Justice Advisory Opinion

A further significant effort to draw the various legal strands together occurred in the mid-1990s with efforts at the World Health Organization (WHO) and the United Nations General Assembly to seek an advisory opinion from the International Court of Justice on the legality of the use, or threat of use, of nuclear weapons. Ultimately, a majority of the court held that the WHO's efforts went beyond its constituted powers.

The court, however, had few qualms about trying to answer the concerns of the General Assembly, because the United Nation held much wider competence on questions regarding peace and security. The Assembly asked: Is the threat or use of nuclear weapons in any circumstance permitted under international law? The court rendered its opinion on July 8, 1996. States opposed to nuclear weapons argued that the use, or threat of use, of nuclear weapons is illegal in itself, any time and anywhere. Three of the fourteen judges on the court agreed. Seven more said that it would "generally" be contrary to the laws of war to use or threaten to use nuclear weapons. The seven added that they were not sure whether such a use "would be lawful or unlawful in an extreme circumstance of self-defence, in which the very survival of a State would be at stake." Four judges, Stephen Schwebel (United States), Sheru Oda (Japan), Gilbert Guillaume (France), and Rosalyn Higgins (United Kingdom), disagreed with both of these positions: They believed that each individual case had to be considered against the relevant standards and that no general rule was possible.

The arguments primarily drew upon the law of armed conflict (humanitarian law); environmental law; human rights law (especially the right to life and the law relating to genocide); and the constitutional documents of the UN and the WHO—the UN Charter and the WHO Constitution. Opponents of nuclear weapons argued that these bodies of law pointed, individually or cumulatively, in the direction of the illegality of nuclear weapons. Instruments such as the Partial Test Ban Treaty (PTBT) of 1963 and the Non-Proliferation Treaty (NPT) of 1968 were said to provide further indications of the aversion of international law to nuclear weaponry. The 1963 treaty bans nuclear weapons tests in the atmosphere, in outer space, and under water. The NPT recognizes that the original five nuclear powers—the United States, Russia, The United Kingdom, France and China—already have the weapons, but it nonetheless tries to keep others from developing them.

The essence of the argument by the nuclear powers was that none of these bodies of law expressly addresses the use of nuclear weapons and that, consequently, there was nothing to prohibit their use, or the threat of their use. Moreover, the NPT, they contended, legitimized the possession and thus potential use of nuclear weapons. The benevolent intentions of the nuclear powers were said to be supported by the "negative security guarantees" given in 1995 by the United States, Russia, the United Kingdom, and France. Essentially, they promised not to use nuclear weapons on a non-nuclear state, unless that state carried out an attack in association or alliance with a nuclear-weapon state. (China made a similar promise, without the exception.) Many developing countries, on the other hand, saw the NPT as discriminatory.

After initial discussion of the court's jurisdiction and of the question itself, the court addressed the arguments that were based on human rights and environmental law. It suggested that human rights arguments are inconclusive where nuclear weapons are concerned because they ultimately send the enquiry to the laws of armed conflict. The court then held that the laws of armed conflict amount to a lex specialis in the present context. In other words, the provisions of the laws of armed conflict would prevail over the more general precepts of human-rights law. The same was true of the environmental arguments. "Respect for the environment is one of the elements that go to assessing whether an action is in conformity with the principles of necessity and proportionality [in the laws of armed conflict]." Similarly, the provisions of the UN Charter on when force is, or is not, lawful do not get to the ultimate conclusion. They have to be read subject to the laws of war—even lawful self defense is subject to the constraints of those rules.

Of particular interest is the discussion of genocide. Some nations had contended that the prohibition contained in the 1948 Convention was a relevant rule of customary law that the court must apply to nuclear weapons. Article I of the Genocide Convention confirms that it applies "in time of peace or in time of war." The court summarized the arguments as follows:

It was maintained before the Court that the number of deaths occasioned by the use of nuclear weapons would be enormous; that the victims could, in certain cases, include persons of a particular national, ethnic, racial, or religious group; that the intention to destroy such groups could be inferred from the fact that the user of the nuclear weapon would have omitted to take account of the well-known effects of the use of such weapons.

According to the court, however, this might sometimes be the case; sometimes not:

The Court would point out in that regard that the prohibition of genocide would be pertinent in this case if the recourse to nuclear weapons did indeed entail the element of intent, towards a group as such, required by the provision quoted above. In the view of the Court, it would only be possible to arrive at such a conclusion after having taken due account of the circumstances specific to each case.

While the Court did not specifically address it, the logic of its argument on genocide must apply also to the invocation of crimes against humanity in the attempt to ban the use of nuclear weapons. Unless the thresholds for a crime against humanity can be shown—an attack on a civilian population, and knowledge of that attack—there is no crime against humanity. Use of a nuclear weapon may, in some ill-defined circumstances, be justified or excused. In others it may be the engine of a crime against humanity. The court saw itself as concerned with international conflict. It could be argued that the most likely kind of case where it would be necessary to concentrate, for purposes of legal analysis, on genocide and crimes against humanity following the use of a nuclear weapon will be in the case of an internal conflict. In that context, the laws of armed conflict are still developing, and there the victims are not in a position to engage in the kind of armed resistance that would bring those laws into play. Thus the court arrived at what it regarded the nub of the debate: the laws of armed conflict.

Opponents of nuclear weapons argued that existing treaty provisions and customary law were broad enough to proscribe nuclear weapons, even though the laws do not say so explicitly and for the most part had been written before nuclear weapons were invented. The laws' relevance could be found, for example, by interpreting treaties (and customary law) that ban the use in armed conflict of items such as poison or asphyxiating substances as also including nuclear weapons. Alternatively, one could look to international customary law (anchored mainly in a series of General Assembly resolutions) specifically proscribing nuclear weapons. Another way to achieve the same end would be to acknowledge that it is impossible to use nuclear and other weapons of mass destruction without contravening the prohibitions of unnecessary suffering, indiscriminate attacks which include civilians as targets, and breaches of the neutrality of non-participants in the conflict. Eleven members of the court thought otherwise, however, stating, "There is in neither customary nor conventional international law any comprehensive and universal prohibition of the threat or use of nuclear weapons as such."

Views of the Court

So far as treaty language banning specific weapons goes, these eleven members did not regard what nuclear weapons do to people as bringing them within prohibitions relating to asphyxiating gases or poisons. Apparently, what radiation does is just incidental to the prime effect of nuclear energy, namely, to blow people to smithereens or to incinerate them. That is different from poisoning or asphyxiating and thus acceptable, or at least not illegal by virtue of the ban on poisons or gases. Moreover, the various treaties on nuclear-free areas and the NPT do not create a general prohibition on the use of nuclear weapons.

Nor did the eleven regard numerous nuclear-specific General Assembly resolutions as sufficient. The series of General Assembly resolutions in question begin with Resolution 1653 of November 24, 1961: the Declaration on the Prohibition of the Use of Nuclear and Thermo-Nuclear Weapons. Adopted by a majority of 55 to 20, with 26 abstentions, it asserted, "Any State using nuclear and thermo-nuclear weapons is to be considered as violating the Charter of the United Nations, as acting contrary to the laws of humanity, and as committing a crime against mankind and civilization." The reference to the laws of humanity evokes the Martens Clause in the preamble to the Fourth Hague Convention of 1907. This clause asserts that, until a more complete code has been attained for the laws of war, "the inhabitants and the belligerents remain under the protection and the rule of the principles of the law of nations, as they result from the usages established among civilized peoples, from the laws of humanity, and the dictates of the public conscience." In the 1981 Declaration on the Prevention of Nuclear Catastrophe, also adopted by a large majority, the Assembly declared, "States and statesmen that resort first to the use of nuclear weapons will be committing the gravest crime against humanity." There is a close historical connection between the Martens Clause and the development of the concept of a crime against humanity, of which genocide is one branch.

Scholars usually assert that customary international law has two elements: consistent practice and a sense of obligation (or opinio juris) concerning that practice. The court acknowledged that although the General Assembly has no general law-making power, its resolutions may have a role in ascertaining customary law:

General Assembly resolutions, even if they are not binding, may sometimes have normative value. They can, in certain circumstances, provide evidence important for establishing the existence of a rule or the emergence of an opinio juris. To establish whether this is true of a given General Assembly resolution, it is necessary to look at its content and the conditions of its adoption; it is also necessary to see whether an opinio juris exists as to its normative character. Or a series of resolutions may show the gradual evolution of the opinio juris required for the establishment of a new rule.

The eleven did not see the failure to use nuclear weapons since 1945 and the practice represented by the line of GA resolutions as enough:

[S]everal of the resolutions under consideration have been adopted with substantial numbers of negative votes and abstentions; thus, although those resolutions are a clear sign of deep concern regarding the problem of nuclear weapons, they still fall short of establishing the existence of an opinio juris on the illegality of the use of such weapons.

The opinion then turns to principles of the law of war, such as unnecessary suffering, indiscriminate targeting, and breaches of neutrality, which the court locates in an overlapping mixture of customary and treaty law. All fourteen judges agreed that these principles apply to nuclear weapons. The opinion even cites statements by the nuclear powers to this effect in the oral pleadings. It is the implication of these principles, which leads to a sharp divergence. "The Court" (in fact seven of the judges, with the tie broken by the unusual rule of the court that gives the President the right to cast a tie-breaking vote in addition to his normal one) offers some cryptic remarks on the topic, summarized at paragraph 105 (2) E of the opinion:

It follows from the above-mentioned requirements that the threat or use of nuclear weapons would generally be contrary to the rules of international law applicable in armed conflict, and in particular the principles and rules of humanitarian law;

However, in view of the current state of international law, and the elements of fact at its disposal, the Court cannot conclude definitively whether the threat or use of nuclear weapons would be lawful or unlawful in an extreme circumstance of self-defense, in which the very survival of a State would be at stake.

The individual opinions of the seven in the "majority" covered a broad spectrum, particularly on the second sub-paragraph of Paragraph E, which dealt with the possible exceptional case—self-defense—when the use of nuclear weapons would not be contrary to international law regarding armed conflict. At one end, some seemed to have doubts about even the validity of the ultimate self-defense exception. At the other, some seemed to accept that there was an in extremis self defense exception.

The seven person dissent comprised two diametrically opposite groups. Judges Christopher Weeramantry, Abdul Koroma, and Mohamed Shahabuddeen voted against the majority finding because they felt that the opinion did not go far enough; Judges Stephen Schwebel, Sheru Oda, Gilbert Guillaume, and Rosalyn Higgins voted against it because they felt the opinion went too far. For Weeramantry, Koroma, and Shahabuddeen, the rules of armed conflict, the specific and the general, proscribe nuclear weapons in all circumstances. No conceivable use of nuclear weapons could comply with the rules. For Schwebel, Oda, Guillaume, and Higgins, the laws of armed conflict apply, but each individual use or threat of use must be considered on its own merits, as would be true of any other weapon that is lawful in itself.

One other inquiry, which the court addressed only inconclusively, related to the nuclear powers' doctrine of deterrence, the argument that the possession of nuclear weapons deterred their use and intimidated non-nuclear nations who might otherwise be tempted to engage in aggression or to use nuclear or other weapons of mass destruction. During the cold war period, it was widely argued that the doctrine of Mutually Assured Destruction (MAD) meant that no leader would dare risk starting a nuclear war in which all might perish. While the court opined that it could not ignore the doctrine, it did not offer a legal characterization of it. Judge Schwebel, in his dissenting opinion, however, seemed to regard the doctrine as supportive of the nuclear powers' position on customary law.

Having split three ways on the crucial issue, the court spoke unanimously regarding a certain matter that was not directly responsive to the question asked. It nonetheless points in the only possible direction now open regarding the issue of nuclear weapons. The presence of this matter in the court reflected widespread frustration that, after nearly thirty years, the promise of Article VI of the 1968 Non-Proliferation Treaty (NPT) had not been fulfilled. Article VI provides that:

Each of the Parties to the Treaty undertakes to pursue negotiations in good faith on effective measures relating to cessation of the nuclear arms race at an early date and to nuclear disarmament, and on a treaty on general and complete disarmament under strict and effective international control.

At the time that the Advisory Opinion was written, 182 countries were parties to the non-proliferation treaty. By the end of 2003, there were 188, but one had claimed to withdraw. The opinion reiterates the Article VI obligation in various ways, hinting that it applies (as customary law) to parties and (the few) non-parties to the treaty alike. There is an obligation both to negotiate in good faith and to achieve a particular result—total nuclear disarmament—as well as to reach the broader goal of general and complete disarmament.

The whole object of the case had been to delegitimize the nuclear bomb. No one doubted that ultimately it would still be necessary to complete the disarmament negotiations. Even total success in the case would not have magically eliminated existing stockpiles. The success of the case in chipping away at the acceptability of nuclear weapons should have made it a little more likely that those negotiations would be completed sooner rather than later.

Nuclear Nonproliferation

The NPT envisaged that conferences would be held at five-year intervals in order to review the operation of the treaty. Concluded in 1968 and in force in 1970, it was initially effective for a period of twenty-five years. In 1995, while the advisory proceedings were pending, the parties agreed that it would continue in force indefinitely. At the review in 2000, a group known as the "New Agenda Coalition" (Brazil, Egypt, Ireland, Mexico, New Zealand, South Africa, and Sweden) spearheaded the effort that resulted in an "unequivocal undertaking by the nuclear-weapons States to accomplish the total elimination of their nuclear arsenals leading to nuclear disarmament to which all States are committed under Article VI."

It is hard to see this vision being realized. In 1997 Costa Rica submitted a Model Nuclear Weapons Convention to the United Nations. Its title says it all: "Convention on the Prohibition of the Development, Testing, Production, Stockpiling, Transfer, Use, and Threat of use of Nuclear Weapons and on Their Elimination." It would lead to progressive prohibition and stringent inspections to ensure compliance. The model has been increasingly refined by nongovernmental groups, such as the Lawyers Committee on Nuclear Policy, but has not captured the imagination of governments. Negotiations proceed glacially in various forums, including the First Committee of the United Nations General Assembly, the sixty-six nation Conference on Disarmament which meets in Geneva and the Assembly's Commission on Disarmament.

Although they have worked toward reducing their arsenals, the nuclear powers seem determined to rely on them in some circumstances, and even to continue research and development. Albeit observing a moratorium on testing, the United States, for example, seeks to develop a "mini-nuke" capable of going after deeply buried weapons of mass destruction. In December 2001, President Bush announced the United States' withdrawal from the 1972 agreement with Russia on the limitation of anti-ballistic missile systems. That agreement complemented the two super-powers' policy of Mutually Assured Destruction (MAD) and was a basic element of their search for deterrence. The 1972 treaty prohibited the parties from putting into place systems capable of defending their entire territories from intercontinental ballistic missiles and from developing, testing, or deploying sea-, air-, space-, or mobile land-based antiballistic missile systems.

Those in favor of withdrawing saw the treaty as an obstacle to developing a comprehensive defense against weapons of mass destruction. Those opposed feared the U.S. government would now embark on an incredibly expensive technological effort, which had no guarantee of success. At the same time, they argued, ending the treaty could result in a new arms race with Russia and even China. Meanwhile, a more pressing danger was posed by terrorists and rogue states with delivery systems other than intercontinental missiles. A relatively small "dirty bomb" or radiological instrument in the hands of terrorists might present a greater danger than a developed bomb, and resources might be better spent in dealing with such dangers.

In December of 2002, the United States issued a new "National Strategy to Combat Weapons of Mass Destruction" which asserts that the United States "reserves the right to respond with overwhelming force—including through resort to all of our options—to the use of WMD against the United States, our forces abroad, friends, and allies." The phrase, "all of our options," clearly includes both conventional and nuclear responses, even in "appropriate cases through preemptive measures." This is perhaps even clearer than a similar statement made earlier in the year in a Nuclear Posture Review. Serious questions have been raised about the compatibility of these moves with the United Nations Charter and with the International Court of Justice's opinion.

Three nations (India, Israel, and Pakistan) have remained resolutely outside the NPT. Another, North Korea, has purported to withdraw. It claims the right under a treaty provision (similar to that the United States invoked in withdrawing from the ABM treaty) that a party "shall in exercising its national sovereignty have the right to withdraw from the Treaty if it decides that extraordinary events, related to the subject matter of this Treaty, have jeopardized the supreme interests of its country." North Korea asserted its security was jeopardized by the United States, which North Korea claimed was threatening a pre-emptive nuclear strike, other military action, and a blockade. North Korea's right to withdraw is hotly debated.

Divisions and Debate

More positive have been developments involving the IAEA's inspection regime. Under the NPT, the IAEA enters into safeguard agreements with non-nuclear weapons states to maintain controls over nuclear material for peaceful activities. Efforts to strengthen that system have been undertaken since 1992, with the discovery of the extent of Iraq's weapons program, notwithstanding the safeguards. These efforts entailed the development of more intrusive reporting and inspection. States are encouraged to accept this by becoming party to an optional protocol, a model of which was developed by the Agency in 1997. Late in 2003, Iran agreed to such a protocol and Libya was about to. The IAEA inspections regime could provide a precedent, along with that developed by the Organization for the Prevention of Chemical Weapons, for a more comprehensive nuclear abolition treaty, along the lines of the model introduced by Costa Rica. Meanwhile, efforts continue to put greater international control over fissile material adaptable to bomb-making.

Shortly after the International Court of Justice rendered its opinion, in September 1996, the United Nations approved the Comprehensive Nuclear Test Ban Treaty (CTBT). Its rationale is expressed succinctly in a preambular paragraph:

The cessation of all nuclear weapon test explosions and all other nuclear explosions, by constraining the development and qualitative improvement of nuclear weapons and ending the development of advanced new types of nuclear weapons, constitutes an effective measure of nuclear disarmament and non-proliferation in all its aspects.

Parties undertake not to carry out any nuclear weapon test explosion or any other nuclear explosion, and to prevent any such nuclear explosion at any place under their control. At the end of 2003, the treaty was not yet in force. While it had over one hundred signatories, by its own terms it cannot come into effect until ratified by forty-four named States that possess nuclear reactors. About three-quarters of them had done so by 2004, including France, the Russian Federation, and the United Kingdom. There were notable holdouts, such as China, the United States (where the treaty was rejected in the Senate), India, Pakistan and North Korea.

An effort to include the use of nuclear weapons as a war crime in the Rome Statute of the International Criminal Court failed in 1998. In a negotiation based on finding consensus, a majority supported it but it was adamantly opposed by the Nuclear Club, and thus failed. The way was left open for re-examination in the future.

Perversely perhaps, the laws of armed conflict regulate the ethics and modalities of killing. They place an absolute ban on certain kinds of weapons, such as exploding bullets below a certain size, dum-dum (expanding) bullets, poison, asphyxiating gases, and bacteriological substances. Use of such weapons is always a war crime, no matter how good the cause. Judge Weeramantry, dissenting in the Nuclear Weapons Case, raised the fundamental question how such modalities can be proscribed, yet permit nuclear weapons to remain lawful:

At least, it would seem passing strange that the expansion within a single soldier of a single bullet is an excessive cruelty which international law has been unable to tolerate since 1899, and that the incineration in one second of a hundred thousand civilians is not. This astonishment would be compounded when that weapon has the capability, through multiple use, of endangering the entire human species and all civilization with it.

One might equally ask whether it is "passing strange" that use of a nuclear weapon is not yet genocide or a crime against humanity as a matter of law. But genocide, as defined in the Genocide Convention, requires a specific mental element, the "intent to destroy, in whole or in part, a national, ethnical, racial, or religious group, as such." It will often be possible to infer such an intent from use of nuclear weapons, but apparently not always. A crime against humanity requires knowledge that what is being done is part of an attack on a civilian population. Again, inferences may be drawn, but some think that may not always be so.

SEE ALSO Hiroshima; Humanitarian Law; International Court of Justice; Weapons of Mass Destruction

BIBLIOGRAPHY

Boisson de Chazournes, Laurence, and Philippe Sands, ed. (1999). International Law, the International Court of Justice and Nuclear Weapons. Cambridge: Cambridge University Press.

Burroughs, John (1997). The Legality of the Threat or Use of Nuclear Weapons: A Guide to the Historic Opinion of the International Court of Justice. Muenster: Lit Verlag.

Clark, Roger S., and Madeleine Sann (1996). The Case against the Bomb: Marshall Islands, Samoa and the Solomon Islands before the International Court of Justice in Advisory Proceedings on the Legality of the Threat or Use of Nuclear Weapons. Camden, N.J.: Rutgers University School of Law.

Criminal Law Forum: An International Journal 7, no. 2 (1996). Contains written submissions to the International Court of Justice, pro and con the legality of Nuclear Weapons, by the United States and Solomon Islands.

International Review of the Red Cross (January–February 1997). No. 316. Special issue on the "Advisory Opinion of the International Court of Justice on the Legality of Nuclear Weapons and International Humanitarian Law."

Lifton, Robert Jay (1990). The Genocidal Mentality: Nazi Holocaust and Nuclear Threat. New York: Basic Books.

Matheson, Michael J. (1997). "The Opinions of the International Court of Justice on the Threat or Use of Nuclear Weapons." American Journal of International Law 91(417).

Moxley, Charles J., Jr. (2000). Nuclear Weapons and International Law in the Post Cold War World. Lanham, Md.: Austin & Winfield.

Nagan, Winston P. (1999). "Nuclear Arsenals, International Lawyers, and the Challenge of the Millenium" Yale Journal of International Law 24(484).

Nanda, Ved, and David Krieger (1998). Nuclear Weapons and the World Court. Ardsley, N.Y.: Transnational Publishers.

Roff, Sue Rabbit (1995). Hotspots: The Legacy of Hiroshima and Nagasaki. New York: Cassell.

Sagan, Carl A. (1990). A Path Where No Man Thought: Nuclear Winter and the End of the Arms Race. New York: Random House.

Schwartz, Stephen I., ed. (1998). Atomic Audit: The Costs and Consequences of U.S. Nuclear Weapons Since 1940. Washington, D.C.: Brookings Institution Press.

Ware, Alyn (2003). "Rule of Force or Rule of Law: Legal Responses to Nuclear Threats from Terrorism, Proliferation, and War" Seattle Journal for Social Justice 2(243).

LEGAL CASES

Australia v. France. International Court of Justice, Nuclear Tests Case, Judgment of 20 December 1974, 1974 I.C.J. Reports 253.

International Court of Justice. The Legality of the Threat or Use of Nuclear Weapons, Advisory Opinion of 8 July 1996, 1996 I.C.J. Reports 226.

New Zealand v. France. International Court of Justice, Nuclear Tests Case, Judgment of 20 December 1974, 1974 I.C.J. Reports 457.

New Zealand v. France. International Court of Justice, Request for an Examination of the Situation in Accordance with Paragraph 63 of the Court's Judgment of 20 December 1974 in the Nuclear Tests Case, Order of 22 September 1995, 1995 I.C.J. Reports 288.

"Shimoda v. State." Japanese Annual of International Law 8, no. 212 (1964).

Roger S. Clark

Nuclear Weapons

views updated Jun 27 2018

NUCLEAR WEAPONS.

ORIGINS AND THE MANHATTAN PROJECT
THE SOVIET BOMB
THE HYDROGEN BOMB
BRITISH AND FRENCH NUCLEAR WEAPONS
MUTUALLY ASSURED DESTRUCTION AND THE CUBAN MISSILE CRISIS
ARMS CONTROL EFFORTS
ABLE ARCHER, GORBACHEV, AND THE END OF THE COLD WAR
BIBLIOGRAPHY

On the morning of 6 August 1945, an American bomber dropped "Little Boy" on the Japanese city of Hiroshima, killing tens of thousands of people and demonstrating to the world the awesome power of a new kind of weapon. Little Boy was an atomic bomb, the product of a secret U.S.-British-Canadian wartime effort known as the Manhattan Project. Over the years that followed, nuclear weapons transformed thinking about war and international politics. They were one of the central facts of the Cold War (1945–1989), shaping the conflict between the superpowers and sparking major crises. Their incredible destructive power threatened to destroy humanity but, simultaneously, formed the basis of international stability and peace.

ORIGINS AND THE MANHATTAN PROJECT

The roots of the Manhattan Project lay in the scientific achievements of the interwar period. In 1932 John Douglas Cockcroft (1897–1967) and Ernest Thompson Sinton Walton (1903–1995) of Cambridge University split an atom of lithium, proving that an atom's nucleus could be broken apart, releasing energy in the process. Building on Cockroft and Walton's work, the Italian physicist Enrico Fermi (1901–1954) showed that the atoms of almost every element could be split via neutron bombardment. This discovery, for which he won the Nobel Prize in 1938, raised the possibility of large-scale nuclear fission. According to this idea, one could start a chain reaction by splitting a few atoms of a radioactive substance (such as uranium), which would release both energy and neutrons, which in turn would split surrounding atoms, releasing still more energy and more neutrons. The result would be a huge amount of energy. Like many other European physicists in the 1930s, Fermi emigrated to the United States in order to escape the rising tide of fascism and anti-Semitism in Europe. Arriving in New York in 1939, and then moving to the University of Chicago, he produced the first controlled nuclear chain reaction in late 1942. The achievement was of major significance to the Manhattan Project, then already under way for more than a year.

In the late 1930s, a number of European émigré scientists worried that the Germans were trying to build an atomic bomb. If this happened, the consequences would be dire. On the basis of these concerns, Albert Einstein (1879–1955)—himself a German-Jewish refugee—wrote to President Franklin Delano Roosevelt (1882–1945) in August 1939, urging him to establish a program to accelerate the research already being done in American universities. Roosevelt agreed. The American government began to award grants for work on nuclear fission. The experiments that resulted formed the core of what became the Manhattan Project, which the president formally established in October 1941 by authorizing the development of atomic weapons.

Although Army General Leslie Richard Groves (1896–1970) was put in charge of the project, the American physicist Julius Robert Oppenheimer (1904–1967) was responsible for its scientific work. Oppenheimer established the project's headquarters at Los Alamos, New Mexico, and assembled a group of top scientists—many of whom had fled Europe—to collaborate on the research. Among those who worked on the project were Fermi, the German Hans Albrecht Bethe (1906–2005), the Hungarian Edward Teller (1908–2003), and the Austrian Victor Weisskopf (1908–2002), all leading experts on nuclear physics. One of the key problems facing the team was how to produce enough uranium-235, the element's fissionable isotope, and plutonium to build a bomb. To accomplish this task, enormous factories were built in Oak Ridge, Tennessee, and Richland, Washington, to extract uranium-235 from uranium ore and produce plutonium by bombarding uranium-238 with protons.

Once Fermi had established that a self-sustaining nuclear chain reaction was in fact possible, the only major obstacle left was how to turn the fissile material into a useful weapon. Two different kinds of bombs were built, one using uranium-235 and the other plutonium. In July 1945, two months after the German surrender, the first test of an atomic bomb took place in the New Mexican desert. Codenamed "Trinity," it produced an even bigger explosion than predicted. Less than a month later, the first atomic weapons used in war were dropped on Hiroshima and Nagasaki.

THE SOVIET BOMB

The American monopoly lasted scarcely four years. Even though the Soviets had been wartime allies, and even though the British and Canadians had been both allies and close collaborators on the Manhattan Project, the U.S. government asserted strict unilateral control over the manufacture of nuclear weapons after the war. Prior to 1945, there had been good reasons not to share any information with the Soviets. Though a partner in the fight against Nazi Germany, the Soviet Union had had an antagonistic relationship with the United States and Western Europe during the interwar period. During the war itself, it was doubtful whether the Soviets would withstand the German invasion. In these circumstances, it would have been foolish to share sensitive information with Moscow, because it could easily have ended up in Nazi hands. After the war, rising U.S.-Soviet tensions gave Washington cause to reconsider any thought of divulging its atomic secrets.

The Soviets made every effort to steal the Manhattan Project's secrets. Thanks to both the high quality of their espionage and sympathy for their cause among certain Western scientists, they were startlingly successful. Their most valuable spy in Los Alamos was Klaus Emil Julius Fuchs (1911–1988), a German-born physicist, devout communist, and longtime Soviet informer. So successful was he that none of his colleagues were aware of his covert activities. His work for Moscow was not discovered until 1950.

The information that he and others passed to the Soviets was of great use in both strategic and scientific terms. At the 1945 Potsdam Conference, the U.S. president Harry S. Truman (1884–1972) informed the Soviet leader Joseph Stalin (1879–1953) that the United States had built a "powerful new weapon." Long aware of the Manhattan Project thanks to the steady stream of information from Los Alamos, the Soviet leader feigned indifference, but was secretly concerned. There is some debate concerning whether Truman attempted—in what is called "atomic diplomacy"—to use the American nuclear monopoly to cow Stalin into concessions regarding eastern Europe. Regardless of whether this was Truman's intention, the Soviets certainly believed that the Americans were trying to frighten them, and they were determined to resist. Rejecting the Baruch Plan, an American proposal to bring atomic weapons under international control, Stalin ordered his scientists to build their own bomb as quickly as possible, and gave them all the resources and intellectual freedom necessary to do so.

They succeeded within four years. In August 1949 the Soviet Union conducted its first nuclear test. The intelligence that the Soviet spies had gathered certainly accelerated the development of the Soviet bomb, but it was by no means essential. The scientists working on the project, led by the physicist Igor Kurchatov (1903–1960) and overseen by Lavrenty Beria (1899–1953), the former head of the NKVD (People's Commissariat of Internal Affairs), were among the world's best and, in all likelihood, they would eventually have succeeded in building the bomb on their own. The infiltration of the Manhattan Project, as dramatic as it was, only saved the Soviets a few years of work. Nevertheless, the 1949 test was a huge surprise to the United States, which had expected to enjoy its monopoly until at least the early 1950s.

THE HYDROGEN BOMB

President Truman responded by ordering the construction of a hydrogen bomb, a fusion weapon that would be several hundred times more powerful than the fission bombs dropped on Japan. The decision was a controversial one among American scientists, many of whom, including Oppenheimer, opposed any further research on such devastating weapons. A hydrogen bomb would be so powerful, they argued, that it could only be used against civilian populations and was therefore inherently a tool of genocide. However, a number of equally eminent scientists supported Truman, including Edward Teller. Teller insisted that the Soviets would build a fusion bomb regardless of what the Americans did. The United States could not afford to be intimidated by the Soviet Union, leaving it no choice but to build a bomb of its own. Along with the hydrogen bomb came the counterintuitive idea that the weapon could be the cornerstone of international peace. If each side could threaten the other with total destruction, then neither would be willing to risk all-out war. This was the kernel of nuclear deterrence, the principle at the heart of Western and Soviet strategy for the duration of the Cold War.

By the time Truman approved the project in January 1950, the Soviets had already made good progress toward their own hydrogen bomb. Unlike their atomic bomb, which was largely based on American plans, Soviet scientists designed this weapon without outside help. Working under the guidance of the brilliant and young Andrei Sakharov (1921–1989), they scored a notable victory, detonating a deliverable bomb in August 1953, more than two years before the Americans. It is noteworthy that the Soviet government never wrestled with the same moral concerns about the hydrogen bomb that the Americans did, because from Moscow's point of view there was little qualitative difference between fission and fusion weapons. The American hydrogen bomb team, with Teller at its head, had tested a hydrogen device ahead of the Soviets, in November 1952 at Eniwetok Atoll in the Pacific. However, this device was so large and unwieldy—weighing more than eighty tons and requiring an enormous refrigerator—that it was of no military use. The test proved that it was possible to build a fusion device, but it took another three years for the Americans to build a usable bomb.

BRITISH AND FRENCH NUCLEAR WEAPONS

By the early 1950s, major changes in both American and Soviet strategy were under way. Nuclear and thermonuclear weapons were so powerful that it was difficult to consider their use in anything but all-out war. But in order to deter one's opponents from even low-level aggression, it was essential to convince them that one was willing to respond with all one's might. This logic undergirded President Dwight Eisenhower's (1890–1969) doctrine of massive retaliation, announced in 1954. One of the key problems with the strategy, however, was persuading American allies of its wisdom.

Right from the end of the Second World War, the United States refused to share either its nuclear research or the weapons themselves with its NATO (North Atlantic Treaty Organization) allies. Washington had pledged to defend these countries against Soviet attack, but insisted on retaining the final say over if and when nuclear weapons would be used. The Western Europeans grew increasingly uncomfortable with this arrangement and disliked their outright reliance on a potentially unreliable ally on such a fundamental issue as national defense. As a result, the British and French insisted on developing their own independent nuclear deterrents.

British scientists had played a major role in the Manhattan Project but lacked access to crucial parts of the bombs' design. More work was needed. There was little debate within the new Labour government of Clement Richard Attlee (1883–1967) about whether it was worthwhile. Foreign Secretary Ernest Bevin (1881–1951), an adamant anticommunist, insisted that, as a great power with a large empire to govern, Britain needed its own nuclear arsenal. However, there was more to the question than just prestige. Wartime bombardment proved how vulnerable the country was to aerial attack, and nuclear weapons were the best way to deter any potential Soviet aggression. For these reasons, in January 1947 the government approved plans to build the bomb. In October 1952, shortly before the Americans tested their first hydrogen device, the British detonated an atomic bomb off the coast of Western Australia.

It took France rather longer to build an A-bomb. In the immediate aftermath of the Second World War, Charles de Gaulle (1890–1970) believed the country needed its own nuclear weapons, but France had fallen so far behind its allies in science and industry during the Nazi occupation that huge obstacles lay in its path. Strong domestic opposition from the Left meant that there was little political will behind the idea. However, after the 1954 defeat at Dien Bien Phu, the French government was determined to rebuild national prestige. The design and production of nuclear weapons was one way to achieve this goal, and the government of Pierre Mendès-France (1907–1982) approved the project. When de Gaulle returned to power in 1958, he gave the idea his full support. France conducted its first successful test in February 1960.

The logic behind the independent French nuclear arsenal, known as the force de frappe, was clear. De Gaulle was adamant that France had to remain a great power. In this light, it was unacceptable to rely on a foreign country—in this case, the United States—on a matter that went to the very heart of national pride. He also shared many of London's fears about the American commitment to his country's defense and believed that, for reasons of both security and prestige, France needed to maintain a certain distance from the United States and NATO on military matters. There was no way to guarantee Washington's full support in the event of Soviet aggression, so France had to be ready to defend itself. Although these were reasonable concerns in theory, they did not hold up well in practice. The force de frappe came into being, but it was an open secret that it would never have been possible without a significant amount of American support. Just as Britain had to abandon its independent nuclear program in 1960 in order to cut costs, France still had to rely on the United States.

MUTUALLY ASSURED DESTRUCTION AND THE CUBAN MISSILE CRISIS

In nuclear weapons' first decade, they existed only in the form of bombs to be dropped from the air. The advances of the 1950s brought new ways to wage nuclear war. Alongside its nuclear research, the Soviet Union had been working on rocket technology with a view to building rockets that could be sent into space. It reached its target in 1957, when it launched a small satellite named Sputnik into orbit. Sputnik itself was not of great strategic value, but the rocket that propelled it was. If the Soviets could send a rocket into space, it followed that they could also hit any point in the world with that same rocket. They now had the ability to strike in the United States with a nuclear-armed missile.

The advent of Inter-Continental Ballistic Missiles (ICBMs) changed the Cold War's nuclear calculus. Prior to 1957, much of the continental United States was safe from atomic attack because it was out of range of even the best Soviet bombers. Sputnik destroyed this immunity. The result was panic in both the United States and Western Europe, and the start of a crash program to develop American missiles. A year later—with the help of German scientists such as Wernher von Braun (1912–1977), who had previously worked for the Nazis—the Americans succeeded, sending their own satellite into orbit. Nevertheless, there were widespread fears in Washington of a "missile gap" between the United States and the USSR, spurred on by the characteristically folksy boast of Nikita Khrushchev (1894–1971) that his country was producing missiles as quickly as it did sausages. Through U-2 surveillance flights over the Soviet Union, the Americans soon learned that the advantage in missile production was in fact theirs.

Regardless, the combination of ICBMs and hydrogen bombs meant that the stakes of the nuclear standoff were as high as they could possibly be. Before the development of the H-bomb, there remained the possibility that nuclear war, however horrible, could remain limited. Now, however, a full-scale war waged with thermonuclear missiles would almost certainly destroy entire countries and might even make the planet uninhabitable. On this basis arose the principle of mutually assured destruction, appropriately known as MAD. Neither the United States nor the USSR stood a chance of surviving, let alone winning, a nuclear war. In theory, this certainty would guarantee peace, because neither side could afford even the smallest risk of hostilities.

The theory came under severe strain during the Cuban missile crisis. The crisis was the single most dangerous moment of the Cold War and proved the risks of the nuclear arms race to both superpowers. The Soviet Union began sending nuclear missiles to Cuba after the Bay of Pigs, a failed American attempt in 1961 to overthrow the island's new socialist leader, Fidel Castro (b. 1926). Khrushchev was eager to support the young Cuban leaders, who reminded him of his own revolutionary past. Moreover, Cuba was an ideal site for the USSR's intermediate-range missiles and a beachhead for spreading communism to Latin America. Khrushchev believed he had the perfect opportunity to retaliate against the recent deployment of intermediate-range American missiles in Turkey and to restore some semblance of a strategic balance between East and West, especially given the USSR's relative lack of ICBMs.

When a U-2 flight discovered the Cuban missiles in October 1962, the American government faced a crisis. It could not tolerate hostile nuclear missiles so close to home, but how could it remove them? The administration of John Fitzgerald Kennedy (1917–1963) considered a number of options, including a full-scale invasion of Cuba, air strikes against the launch sites, and a naval blockade to prevent further missile shipments from getting through. The slightest provocation could trigger a nuclear attack on the United States, which would spark full retaliation and end only in mutual annihilation. Kennedy opted for the blockade. The first Soviet ship to encounter the blockade turned around, opening the door to a settlement. After secret negotiations, Khrushchev agreed to pull the missiles out of Cuba if the United States would remove its missiles from Turkey. Kennedy agreed, and the crisis came to an end.

ARMS CONTROL EFFORTS

The Cuban missile crisis was the clearest indication yet of how easy it would be to tumble into a catastrophic nuclear war. Some kind of arms control was necessary to reduce this danger. The first step in this direction came the year after the crisis, with the signing of the Limited Test Ban Treaty (LTBT). According to its provisions, signatories pledged to stop atmospheric tests of nuclear weapons in order to restrict their development and end the provocations that typically followed each test. All future tests had to happen underground.

The agreement was a modest but signal step, the first attempt to contain the use of nuclear weapons. In August 1963, the Americans, Soviets, and British signed the treaty. Insisting as ever on its independence, the only other nuclear power, France, refused to join them. Two months later, the superpowers also agreed to ban nuclear weapons from space.

Further progress was made in the late 1960s and into the 1970s, an era of calm and stability compared with the early Cold War years. Following the LTBT, the next major milestone was the Nonproliferation Treaty (NPT), concluded in 1968. Despite another French refusal to participate, the Americans, Soviets, and British brought the negotiations to a successful conclusion. The treaty committed them to prevent the spread of nuclear weapons to other countries. The higher the number of states that had nuclear weapons, the reasoning went, the greater the danger of their use in a conflict. Prestige was also on the line, since if the nuclear club could be kept small, the value of membership would remain commensurately high. Indeed, the three signatories gave no serious thought to reducing their stockpiles. It was enough to prevent others from building their own.

Into the 1970s, the superpowers refused to cut back the numbers of their missiles and weapons. They instead agreed to restrict their growth. SALT—the Strategic Arms Limitation Talks—was the result. Starry-eyed idealism motivated neither superpower. Each had pragmatic goals: the Russians some relief from the arms race at a time of economic trouble and the Americans a strategic advantage and fodder for President Richard Nixon's (1913–1994) reelection campaign. Negotiated both by arms control experts in Helsinki and Vienna and between the U.S. national security advisor Henry Kissinger (b. 1923) and the Soviet ambassador Anatoly Dobrynin—the so-called backchannel—the SALT agreements placed limits on antiballistic missile defenses and froze the construction of missile launchers for five years. These results were not overwhelming, but both sides were optimistic that they had built a solid foundation for future progress.

Nixon's successor, Gerald Ford (b. 1913), attempted to maintain the momentum that SALT had generated. In 1974 he and Leonid Brezhnev (1906–1982) agreed on a set of guidelines for the next round of arms control talks, known as SALT II. The goal was to pick up where the first agreement had left off, but the United States refused to discuss its nuclear forces in Europe, a key area of interest for the Soviets. Brezhnev, determined not to let the talks collapse, pressured his domestic critics—especially within the military—to accept the American terms. He succeeded, and negotiators agreed to a maximum of 2,250 missile launchers each. But Ford's successor, Jimmy Carter (b. 1924), insisted on cuts that were even more radical and less acceptable to the skeptics within the Soviet government. Brezhnev and the Soviet military were furious. It took until 1979 to repair the damage and reach a new agreement.

Although a deal was reached, it was never implemented. The United States Senate insisted that the agreement was unverifiable because there was no way to ensure that the Soviets would hold up their end of the bargain. Moreover, the broader process of détente collapsed in the wake of the Soviet invasion of Afghanistan in 1979 and the election of Ronald Reagan (1911–2004) in 1980. Hopes for arms control fell apart as Cold War tensions returned to levels unseen since the early 1960s.

ABLE ARCHER, GORBACHEV, AND THE END OF THE COLD WAR

Reagan came into office on a strong anticommunist platform. Rejecting any further arms control talks, he massively increased the American military budget. Part of this renewed toughness was the deployment of new nuclear missiles—Cruise and Pershing 2—in Western Europe as a response to the Soviets' recent installation of the new SS-20 missiles. Protestors organized massive rallies to oppose the move, but failed to stop it.

In this context of heightened tension, NATO staged a war game, dubbed "Able Archer," in late 1983. The Soviet military detected the exercises but, already on edge because of Reagan's bellicose rhetoric, mistook them for the prelude to a Western attack. Moscow put its forces on alert, ready to respond with nuclear weapons if necessary. The danger was not as acute as it had been during the Cuban missile crises, but both sides still approached the brink of nuclear war without ever intending to. The episode reminded Washington and Moscow of the wisdom of returning to the negotiating table.

The rise to power in 1985 of Mikhail Gorbachev (b. 1931) provided the opportunity for new arms control talks. In 1986 Gorbachev publicly proposed a plan to eliminate all nuclear weapons by the year 2000. The proposal was bold and unprecedented, and it caught the Reagan administration's attention. When the two leaders met at Reykjavik later that year, Gorbachev renewed his offer, and Reagan responded enthusiastically. However, the Soviet leader insisted that the Americans would have to abandon their Strategic Defense Initiative—a research program, popularly known as "star wars," to build a shield against ballistic missiles—before any deal could be reached. Reagan refused. The summit ended without agreement.

This was only a temporary setback. Both sides were now considering massive cuts to their nuclear arsenals, a situation that could not even have been contemplated five years earlier. They renewed their efforts, and reached a deal at the Washington summit in December 1987, agreeing to eliminate their nuclear weapons stationed in Europe. The Intermediate-Range Nuclear Forces (INF) treaty came into force in June 1988. This was the beginning of the end of the nuclear standoff that had lasted almost forty years.

The next major milestone came after the fall of the Berlin Wall in 1989. In July 1991 the United States and the USSR signed the Strategic Arms Reduction Treaty (START), the culmination of negotiations that had begun while Reagan was still in office. The first major commitment to reduce nuclear stockpiles since the invention of the atomic bomb, it required both sides to cut their arsenals by almost a third. It was a major accomplishment and proof that the Cold War was effectively over. The USSR itself collapsed in December, devolving the treaty's obligations onto its successor states. President George Herbert Walker Bush (b. 1924) and his Russian counterpart Boris Yeltsin (b. 1931) signed START II in January 1993. The treaty expanded on START I, radically restricting the possession of nuclear weapons and delivery systems to between three thousand and thirty-five hundred on each side. After ratification by the U.S. Senate in 1996 and the Russian Duma in 2000, it was superseded by the 2002 Strategic Offensive Reduction Treaty (SORT), which capped each side's nuclear arsenal between seventeen hundred and twenty-two hundred warheads.

Nuclear weapons continue to be a major issue in international politics, but the focus has shifted away from the major powers. There is increasing concern, particularly in the West, about the proliferation of nuclear weapons both to smaller states and to nonstate actors, especially terrorist networks. The Indian subcontinent has been relatively stable since the surprise Indian and Pakistani nuclear tests of 1998, but in the early twenty-first century there have been worries that North Korea and Iran were developing weapons of their own. In a similar vein, the American government cited Iraq's alleged nuclear weapons as a justification for the invasion of that country in 2003. Given the number of weapons in existence worldwide—and the availability of so-called suitcase bombs—it is impossible to stop the spread of nuclear technology completely. The threat of nuclear proliferation is one of the severest tests facing the international system in the early twenty-first century. Deterrence provided the basis for stability during the Cold War, but it remains to be seen whether any similar idea can be found to avert the use of nuclear weapons in years to come.

See alsoCold War; Cuban Missile Crisis; Disarmament; Potsdam Conference.

BIBLIOGRAPHY

Beach, Sir Hugh, and Nadine Gurr. Flattering the Passions; or, The Bomb and Britain's Bid for a World Role. London, 1999.

Carlisle, Rodney, ed. Encyclopedia of the Atomic Age. New York, 2001.

Craig, Campbell. Destroying the Village: Eisenhower and Thermonuclear War. New York, 1998.

Fursenko, Aleksandr, and Timothy Naftali. One Hell of a Gamble: Khrushchev, Castro, and Kennedy, 1958–1964. New York, 1997.

Holloway, David. Stalin and the Bomb: The Soviet Union and Atomic Energy, 1939–1956. New Haven, Conn., 1994.

Rhodes, Richard. The Making of the Atomic Bomb. New York, 1986.

Michael D. J. Morgan

Nuclear Weapons

views updated Jun 08 2018

Nuclear Weapons

LARRY GILMAN/

K. LEE LERNER/

DEAN ALLEN HAYCOCK

Nuclear weapons are explosive devices that utilize the processes of fission and fusion to release nuclear energy. An individual nuclear device may have an explosive force equivalent to millions of tons (megatons) of trinitrotoluene (TNT, the chemical explosive traditionally used for such comparisons), more than enough to completely destroy a large city. The destructive power of nuclear weapons derives from the core of the atom, the nucleus. One type of nuclear weapon, the fission bomb, uses the energy released when nuclei of heavy elements, such as plutonium, fission or split apart. A second even more powerful type of nuclear weapon, the fusion or hydrogen bomb, uses the energy released when nuclei of hydrogen are forced to fuse (join together).

Nuclear devices have been fashioned into weapons of many shapes with many purposes. Bombs can be dropped from airplanes; warheads can be delivered by missiles launched from land, air, or sea; artillery shells can be fired from cannons; mines can be placed on the land and in the sea. Some nuclear weapons are small enough to destroy only a portion of a battlefield; others, as already mentioned, are large enough to destroy entire cities and more.

Unlike chemical explosives, nuclear weapons have had no peacetime uses, although in the 1950s the U.S. government briefly considered using them to blast artificial harbors in the Alaskan coastline. They are possessed by a number of nations, including the United States, France, Great Britain, China, India, Israel, Pakistan, and the Russian Federation along with several former Soviet Republics. Iran and North Korea, among other nations, are interested in building them. Since nuclear weapons were invented during World War II, they have been used only twice, both times against cities in Japan by the United States.

Development of nuclear weapons. German physicist Albert Einstein (18791955) did not know it at the time, but when he published his Special Theory of Relativity in 1905 he provided the world with the basic information needed to build nuclear weapons. Einstein said that the amount of matter of an object (i.e., its mass) is equivalent to a specific amount of energy. The exact amount of energy in an object equals its mass multiplied by the square of the speed of light. The speed of light is large186,282 miles per second (300,000 km/sec)so even a small piece of matter contains a vast amount of energy. A baseball-size sample of uranium-235, for example, can explode with as much energy as 20,000 tons of TNTand this involves the conversion of only a tiny fraction of the uranium's mass into energy. One pound of explosive material in a fission weapon is approximately 100,000 times as powerful as one pound of TNT.

As World War II approached, two German chemists, Fritz Strassmann (19021980) and Otto Hahn (18791968), pointed a stream of neutrons at a sample of uranium and succeeded in splitting the nuclei of some of its atoms. This splitting of nuclei is termed nuclear fission. The energy released through nuclear fission was the source of power for the first atomic bomb, which was built in the United States by a large team of scientists led by U.S. physicist J. Oppenheimer (19041967). This secret research and development program was termed the Manhattan Project.

The first atomic bomb was detonated in a test at Alamogordo, New Mexico, on July 16, 1945. Three weeks later, on August 6, a bomber named Enola Gay dropped a four-ton atomic bomb containing 12 lb (5.4 kg) of uranium-235 on the Japanese city of Hiroshima. Seventy thousand people died as a direct result of the blast. Within two months, nearly twice that many were dead from blast injuries and radiation. Three days later, on August 9, a bomb containing several pounds of plutonium was dropped on Nagasaki. Thirty thousand people died in the seconds following the explosion, and more later. The Japanese surrendered the next day, ending World War II.

These first nuclear weapons were atomic bombs or A-bombs. They depended on the energy produced by nuclear fission for their destructive power. However, scientists like U.S. physicist Edward Teller (1908) knew even before the first atomic bomb exploded that the fission weapons could be used to create an even more powerful explosive, now called a thermonuclear device, hydrogen bomb, or H-bomb. This weapon gets it power from the energy released when atoms of the hydrogen isotopes deuterium or tritium are forced together, a process called nuclear fusion. Starting a nuclear fusion reaction is even more complicated than setting off a fission atomic bomb; it requires such heat to initiate it that a fission bomb is used as a detonator to explode the fusion bomb. The United States tested the first hydrogen bomb on November 1, 1952. It exploded with the force of 10.4 megatons (millions of tons of TNT equivalent). Three years later, the Soviet Union exploded a similar device.

For the next 40 years, the United States, with its allies, and the former Soviet Union, with its allies, raced to build more nuclear weapons, with each side producing tens of thousands. The end of the cold war and the breakup of the Soviet Union in the early 1990s led to the elimination of a significant number of nuclear weapons; however, the U.S. and Russia still possess many thousands of nuclear weapons.

The physics and mechanics of nuclear weapons. Conventional, chemical explosives get their power from the rapid rearrangement of chemical bonds, the links between atoms made by sharing electrons. In chemical explosives, atoms dissociate from other atoms and form new associations; this releases energy, but the atoms themselves do not change. Nuclear weapons are based on an entirely different principle. They derive their explosive power from changes in the structure of the atom itself, specifically, in the core of the atom, its nucleus.

Atomic bombs use the energy released when nuclei of heavy elements split apart or fission. Uranium and plutonium are the two elements that can be used as fuel for this type of weapon. When nuclei of these atoms are struck with rapidly moving neutrons, they are broken into two pieces nearly equal in size. They also release more neutrons, which split more nuclei. This is called a chain reaction. If enough atomic nuclei split they will release enough neutrons to ensure that all the nuclei of all the atoms in a sample will be split. Enormous amounts of energy are then released in a fraction of a second. This release of energy is the power behind the atomic bomb.

Uranium and plutonium are termed fissile materials because they can support a fission chain reaction if enough material is concentrated in one place. Too small a sample would not generate enough neutrons to keep the fission process going; for example, a one-pound (.45-kg) sample of uranium-235, a sample about the size of a ping-pong ball, is not large enough to support a chain reaction. The atomic bombs used in World War II proved that 12 or so pounds (about 5.5 kg) of fissile material, larger than a ping-pong ball but still small enough to fit into your hand, is enough to maintain a chain reaction. The smallest amount of material that can support a chain reaction is called the critical mass.

The instant enough bomb material is gathered together into a critical mass, a chain reaction begins. (At higher density, less mass is required.) This means that fissile material cannot be assembled in a critical mass until it is meant to explode. Therefore, the sample of uranium or plutonium in an atomic bomb is separated into several pieces, each of which is below critical mass. To set the bomb off, the separated pieces of bomb material are rammed together to create a critical mass. One design for creating a critical mass involves firing a subcritical "bullet" of fissile material into a subcritical "target" of fissile material. Together, the bullet and the target create a critical mass that starts a chain reaction leading to a nuclear explosion.

A different design was used to detonate the bomb dropped on Nagasaki. Plutonium was stored in one large but subcritical mass. It was compressed to a critical density by means of surrounding chemical explosives. When the chemical explosive detonated, the blast forced the bomb material into a density that reached criticality. In either type of design, once criticality is reached the explosion follows in a millionth of a second.

In order for nuclear fission to occur, a bomb must use heavy atoms for fuel. Heavy atoms have many nucleonsneutrons and protonsin their nuclei. When these heavy nuclei split apart they release energy (and neutrons, which may cause nearby heavy nuclei to split apart also). Another more powerful type of nuclear weapon uses forms of hydrogen as fuel. Hydrogen has few subatomic particles in its nucleiusually only a proton, but the isotope deuterium has a proton plus a neutron, while the isotope tritium has a proton plus two neutrons. Instead of being split apart, these light atomic nuclei are forced together in high-speed collisions, a process called nuclear fusion. Energy is released when hydrogen nuclei fuse, forming helium. Fusion only occurs at temperatures of millions of degrees, such as exist in the hearts of stars. (The sun and other stars generate their energy primarily by fusing hydrogen into helium.) On Earth only an atomic bomb can raise kilograms of material to such a temperature, which is why atomic bombs are used as detonators for hydrogen fusion bombs.

Because hydrogen is lighter than uranium, more hydrogen atoms fit into a sample of the same weight. Thus, even though one fusion reaction releases less energy than one fission reaction, more hydrogen than uranium atoms can be packed into a nuclear weapon and many more fusion reactions can take place in the weapon than fission reactions can take place in a fission bomb. Fusion weapons, therefore, produce bigger explosions than fission weapons of the same physical bulk.

By 1954, a new feature had been added to the hydrogen bomb to create an even more dangerous weapon. Like earlier hydrogen bombs, this weapon was detonated with the explosion of an atomic or fission weapon. This raised temperatures enough to cause the hydrogen atoms in the bomb to fuse and explode like a regular hydrogen bomb. The designers also enclosed this new bomb in a shell of uranium-238. Neutrons released from the fusion of hydrogen caused the uranium-238 in the surrounding jacket to undergo fission, adding to the power of the blast. This new device was, in effect, a fission-fusion-fission bomb.

The power or "yield" of a nuclear weapon is expressed in terms of how much TNT would be required to equal the weapon's blast. Units of kilotons (thousands of tons) and megatons (millions of tons) of TNT are used to describe nuclear blasts.

Effects of nuclear weapons. Nuclear weapons produce two important effects that are also produced by conventional, chemical explosives: they release heat and generate shock waves, or pressure fronts of compressed air that smash objects in their paths. The heat released in a nuclear explosion creates a sphere of burning, glowing gas that can range from hundreds of feet to miles in diameter, depending on the power of the bomb. This fireball emits a flash of heat that travels outward from the site of the explosion (ground zero), the area directly under the explosion. This heat can cause second degree burns to bare human flesh miles away from the blast site if the bomb is large enough. (Although this heat can start fires, it seems that much of the fire damage in Hiroshima and Nagasaki following the nuclear explosions resulted from damaged electrical, fuel, gas, and other systems following physical damage caused by the shock or blast wave that accompanied the explosion.)

The shock wave produced when a nuclear weapon explodes creates a front of moving air more powerful than any produced by a natural storm. Destructive winds follow the front of displaced air, causing more damage to objects in their path. Many nuclear weapons are designed to be detonated high above their targets to take advantage of this shock effect. The more powerful the bomb, the higher in the sky it will be detonated. The fission bombs dropped on Japan (Hiroshima, 13.5 kilotons; Nagasaki, 22 kilotons) exploded between 1,500 and 2,000 feet (458610 m) above their targets. A bomb with the power of 10 megatons is capable of destroying most houses within a distance of more than 10 miles from the blast site.

Unlike conventional explosives, nuclear devices can also release significant amounts of radioactivity and pulses of electromagnetic energy. Radioactivity is the release of fast particles and high-energy photons from unstable atomic nuclei. Besides the greater explosive power of nuclear weapons, radiation is the primary feature that most clearly distinguishes chemical from nuclear explosions. Radiation can kill outright at high doses and cause illnesses, including cancer, at lower doses. The initial burst of radiation during a nuclear explosion is made up of X rays, gamma rays, and neutrons. The energy of this radiation is so high that it can often penetrate buildings. Radioactive materials then contaminate the explosion site and often enter the atmosphere where they can travel thousands of miles before falling back to earth. This source of radiation is called radioactive fallout. Radioactive fallout can harm living things for years following a nuclear explosion. Fission bombs and fission-fusion-fission bombs produce more fallout than hydrogen bombs because the fusion of hydrogen atoms generates less radioactive byproducts than does fission of uranium or plutonium.

Electromagnetic pulses (EMPs) are also produced by nuclear weapons that are exploded at high altitudes, and are caused by the interaction of radiation from the explosion with electrons in the atmosphere and with the Earth's magnetic field. EMPs are essentially powerful radio waves that can destroy many electronic circuits.

The effects of fires and destruction following a largescale nuclear war could even change the climate of the planet. In 1983 a group of scientists, including U.S. astronomer Carl Sagan (19341996), published the "nuclear winter" theory, which suggested that particles of smoke and dust produced by fires caused by many nuclear explosions would, for a time, block the Sun's rays from reaching the surface of Earth. This, in turn, would reduce temperatures and change wind patterns and ocean currents. These climatic changes, according to the theory, could destroy crops and lead to the death by famine of many more animals and humans than were killed outright by nuclear explosions. Some scientists have challenged these predictions, but others, including some United States government agencies, support them. On the other hand, there is no controversy about whether a large-scale nuclear war could kill hundreds of millions of people and imperil the future of modern civilization, even apart from nuclear winter effects.

Modern nuclear weapons. Today nuclear weapons are built in many sizes and shapes not available in the 1940s and 1950s, and are designed for use against many different types of military and civilian targets. Some weapons are less powerful than 1,000 tons of TNT, while others have the explosive force of millions of tons of TNT. Small nuclear shells can be fired from cannons. Nuclear warheads mounted on missiles can be launched from land-based silos, ships, submarines, trains, and large wheeled vehicles. Several warheads can be fitted into one missile and directed to different targets in the same geographic area upon reentry into the Earth's atmosphere. These multiple independently-targeted reentry vehicles (MIRVs) can release 10 or so individual nuclear warheads far above their targets, making enemy interception more difficult and increasing the deadliness of each individual missile.

In general, nuclear weapons with "low" yields (in the kiloton, rather than the megaton, range) are termed "tactical," and are designed to be used in battle situations against specific military targets, such as a concentration of enemy troops or tanks, a naval vessel, or the like. These weapons are termed tactical because the word tactics, in military jargon, refers to the relatively small-scale maneuvers undertaken to win particular battles. Larger nuclear weapons are classed as "strategic," because the word strategy, again in military jargon, refers to the large-scale maneuvers undertaken to win whole wars. Strategic nuclear weapons are targeted mostly at cities and at other nuclear weapons, and are generally designed to be dropped by bombers or launched on ballistic missiles; tactical nuclear weapons are delivered by smaller devices over shorter distances. However, one nation's "tactical" warhead may be another's "strategic" warhead: Russia, for example, maintains that U.S. tactical warheads in Western Europe are in fact strategic warheads, because they can strike targets inside Russia itself, while Russian "tactical" warheads in the same arena cannot strike the U.S. heartland.

In the summer of 2002, the George W. Bush administration sought and received permission from Congress to design a new class of nuclear weapons: "mini-nukes" are relatively low-yield tactical nuclear weapons for use against underground bunkers and other small battlefield targets. Also in 2002, the U.S. militaryaccording to a secret Pentagon document leaked to the pressdrew up an official set of contingency plans for attacking seven countries with nuclear weapons (China, Russia, Iraq, North Korea, Iran, Libya and Syria). Advocates of these new weapons point to the uniquely powerful, compact "punch" that can be delivered by a nuclear weapon; critics argue that even a small nuclear weapon may cause many civilian casualties, and, more important, that actual use of a nuclear weapon of any size would break the taboo on such use that has held since the end of World War II, making the use of larger, more destructive nuclear weapons more likely in future conflicts. Some analysts stressed that the Pentagon's explicit willingness to use nuclear weapons in a "first-use" fashion, that is, in response to "unexpected military situations" not involving attack on U.S. forces by nuclear weapons, or to use them on targets (e.g., deep bunkers) resistant to conventional explosives signaled a major shift in United States nuclear use doctrine.

Even the ability of nuclear weapons to release radioactivity has been exploited to create different types of weapons. "Clean bombs" are weapons designed to produce as little radioactive fallout as possible. A hydrogen bomb without a uranium jacket would produce relatively little radioactive contamination, for example. A "dirty bomb" could just as easily be built, using materials that contribute to radioactive fallout. Such weapons could also be detonated near Earth's surface to increase the amount of material that could contribute to radioactive fallout. "Neutron" bombs have been designed to shower battle fields with deadly neutrons that can penetrate buildings and armored vehicles without destroying them. Any people exposed to the neutrons, however, would die. (Neutron bombs also destroy with blast effects, but their deadly radiation zones extend far beyond the site of their explosions).

The United States and Russia signed a Strategic Arms Reduction Treaty in 1993 to eliminate two thirds of their nuclear warheads in 10 years. By 1995, nearly 2,500 nuclear warheads had been removed from bombers and missiles in the two countries, according to U.S. government officials. ("Elimination," in this context, does not necessarily mean dismantlement; many of the weapons that have been "eliminated" by the treaty have been put in storage.) Although thousands of nuclear weapons still remain in the hands of many different governments, especially those of the U.S. and the Russian Federation, recent diplomatic trends have at least helped to lower the number of nuclear weapons in the world. This has caused many people to assume that the danger of nuclear weapons evaporated with the end of the Cold War.

However, the number of nations possessing nuclear weapons continues to increase, and the possibility of nuclear weapons being used against human beings for the first time since World War II may be larger than ever. In May 1995, more than 170 members of the United Nations agreed to permanently extend the Nuclear Non-Proliferation Treaty, first signed in 1960. Under the terms of the treaty, the five major countries with nuclear weaponsthe United States, Britain, France, Russia, and Chinaagreed to commit themselves to eliminating their arsenals as an "ultimate" goal. The other 165 signatory nations agree not to acquire nuclear weapons. Israel, which is believed to possess nuclear weapons (but officially denies doing so), did not sign the treaty. Two other nuclear powers also refused to renounce nuclear weapons: India and Pakistan, each of which probably possess several dozen nuclear weapons, have fought a number of border wars in recent decades, and in 2002 came close, as many observers thought, to fighting a nuclear war. As of 2003, North Korea had reactivated its nuclear-weapons-material production facilities and was engaged in a tense diplomatic standoff with the United States, which insisted that North Korea abandon its nuclear-weapons program.

FURTHER READING:

BOOKS:

Rhodes, Richard. Dark Sun: The Making of the Hydrogen Bomb (Sloan Technology Series). New York: Simon & Schuster, 1995.

Sagan, Scott D. and Kenneth N. Waltz. The Spread of Nuclear Weapons: A Debate Renewed, 2nd ed. W. W. Norton & Co., 2003.

Walmer, Max. An Illustrated Guide to Strategic Weapons. New York: Prentice Hall Press, 1988.

ELECTRONIC

"U.S. Has Nuclear Hit List." BBC News. March 2, 2002. <http://news.bbc.co.uk/2/hi/americas/1864173.stm> (Feb. 26, 2003).

SEE ALSO

Arms Control, United States Bureau
Iranian Nuclear Programs
Manhattan Project
North Korean Nuclear Weapons Programs
Nuclear Detection Devices
Russian Nuclear Materials, Security Issues
World War II

Nuclear Weapons

views updated Jun 11 2018

Nuclear weapons

Nuclear weapons are explosive devices that release nuclear energy . An individual nuclear device may have an explosive force equivalent to millions of tons (megatons) of trinitrotoluene (TNT, the chemical explosive traditionally used for such comparisons), and is more than enough to inflict devastating physical damage to a city.

The destructive power of nuclear weapons derives from the core of the atom, the nucleus. One type of nuclear weapon, the fission bomb, uses the energy released when nuclei of heavy elements such as plutonium fission (split apart). A second even more powerful type of nuclear weapon, the fusion or hydrogen bomb, uses the energy released when nuclei of hydrogen are united (fused together).

Nuclear devices have been fashioned into weapons of many shapes with many purposes. Bombs can be dropped from airplanes; warheads can be delivered by missiles launched from land, air, or sea; artillery shells can be fired from cannon; mines can be placed on the land and in the sea. Some nuclear weapons are small enough to destroy only a portion of a battlefield; others, as already mentioned, are large enough to destroy entire cities or major targets.

Unlike chemical explosives , nuclear weapons have had no peacetime uses. In the 1950s the U.S. government briefly considered using nuclear weapons to blast artificial harbors in the Alaskan coastline but eventually discarded the idea. As of February 2003, nuclear weapons are possessed by a number of nations, including the United States, France, Great Britain, China, India, Israel, Pakistan, and the Russian Federation and several other former Soviet Republics. In February 2003, the United States Central Intelligence Agency released reports that confirm the agency's assertion that North Korea posed one or two nuclear weapons. Iran, Iraq and other nations, have actively attempted to buy components to develop nuclear weapons. Intelligence agencies asserted that, as of February 2003, neither Iran nor Iraq had an operational nuclear weapon but that both countries had nuclear programs capable of producing such weapons.

Since their invention during World War II, nuclear weapons have been used only twice, both times against cities in Japan by the United States. This use of nuclear weapons ended WW II and although horrific in their effect, most historians and analysts assert that the use of the weapons ultimately saved tens of thousands of lives and billions of dollars of destruction by forcing a Japanese surrender and thus making it unnecessary for the United States and other allies to invade the Japanese homeland.


Development of nuclear weapons

German physicist Albert Einstein (1879–1955) did not know it at the time, but when he published his Special Theory of Relativity in 1905 he provided the world with the basic information needed to build nuclear weapons. One aspect of Einstin's work (embodied in the famous equation E= mc2) stated that the amount of matter of an object (i.e., its mass ) is equivalent to a specific amount of energy. The exact amount of energy in an object equals its mass multiplied by the square of the speed of light . The speed of light is large—186,282 miles per second (300,000 km/sec)—so even a small piece of matter contains a vast amount of energy. A baseball-size sample of uranium-235, for example, can explode with as much energy as 20,000 tons of TNT—and this involves the conversion of only a tiny fraction of the uranium's mass into energy. One pound of explosive material in a fission weapon is approximately 100,000 times as powerful as one pound of TNT.

As World War II approached, two German chemists, Fritz Strassmann (1902–1980) and Otto Hahn (1879–1968), pointed a stream of neutrons at a sample of uranium and succeeded in splitting the nuclei of some of its atoms . This splitting of nuclei is termed nuclear fission . The energy released through nuclear fission was the source of power for the first atomic bomb, which was built in the United States by a large team of scientists lead by U.S. physicist J. Oppenheimer (1904–1967). This secret research and development program was termed the Manhattan Project.

The first atomic bomb was detonated in Alamogordo, New Mexico, on July 16, 1945. Three weeks later, on August 6, a United States bomber, the Enola Gay, dropped a four-ton atomic bomb containing 12 lb (5.4 kg) of uranium-235 on the Japanese city of Hiroshima. Seventy thousand people died as a direct result of the blast. Within two months, nearly twice that many were dead from blast injuries and radiation . Three days later, on August 9, a bomb containing several pounds of plutonium was dropped on Nagasaki. Thirty thousand people died in the seconds following the explosion, and more later. The Japanese surrendered the next day, ending World War II.

These first nuclear weapons were atomic bombs or A-bombs. They depended on the energy produced by nuclear fission for their destructive power. However, scientists like U.S. physicist Edward Teller (1908–) knew even before the first atomic bomb exploded that the fission weapons could be used to create an even more powerful explosive, now called a thermonuclear device, hydrogen bomb, or H-bomb. This weapon gets it power from the energy released when atoms of the hydrogen isotopes deuterium or tritium are forced together, a process called nuclear fusion . Starting a nuclear fusion reaction is even more complicated than setting off a fission atomic bomb; it requires such heat to initiate it that a fission bomb is used as a detonator to explode the fusion bomb. The United States tested its first hydrogen bomb on November 1, 1952. It exploded with the force of 10.4 megatons (millions of tons of TNT equivalent). Three years later, the Soviet Union exploded a similar device.

For the next 40 years, the United States, with its allies, and the former Soviet Union, with its allies, raced to build more nuclear weapons. Each side produced tens of thousands of nuclear weapons. The end of the cold war and the breakup of the Soviet Union in the early 1990s led to a significant decrease in the numbers of nuclear weapons in the world; however, the U.S. and Russia still possess many thousands of nuclear weapons.


How nuclear weapons work

Conventional, chemical explosives get their power from the rapid rearrangement of chemical bonds, the links between atoms made by sharing electrons. In chemical explosives, atoms dissociate from other atoms and form new associations; this releases energy, but the atoms themselves do not change. Nuclear weapons are based on an entirely different principle. They derive their explosive power from changes in the structure of the atom itself, specifically, in the core of the atom, its nucleus.

Atomic bombs use the energy released when nuclei of heavy elements split apart or fission. Uranium and plutonium are the two elements that can be used as fuel for this type of weapon. When nuclei of these atoms are struck with rapidly moving neutrons, they are broken into two nearly equal size pieces. They also release more neutrons, which split more nuclei. This is called a chain reaction. If enough atomic nuclei split they will release enough neutrons to ensure that all the nuclei of all the atoms in a sample will be split. Enormous amounts of energy are then released in a fraction of a second. This release of energy is the power behind the atomic bomb.

Uranium and plutonium are termed fissile materials because they can support a fission chain reaction if enough material is concentrated in one place. Too small a sample would not generate enough neutrons to keep the fission process going; for example, a 1-lb (.45-kg) sample of uranium-235, a sample about the size of a ping-pong ball, is not large enough to support a chain reaction. The atomic bombs used in World War II proved that 12 or so pounds (about 5.5 kg) of fissile material, larger than a ping-pong ball but still small enough to fit into a hand, is enough to maintain a chain reaction. The smallest amount of material that can support a chain reaction is called the critical mass.

The instant enough bomb material is gathered together into a critical mass, a chain reaction begins. (At higher density , less mass is required.) This means that fissile material cannot be assembled in a critical mass until it is meant to explode. Therefore, the sample of uranium or plutonium in an atomic bomb is separated into several pieces, each of which is below critical mass. To set the bomb off, the separated pieces of bomb material are rammed together to create a critical mass. One design for creating a critical mass involves firing a subcritical "bullet" of fissile material into a subcritical "target" of fissile material. Together, the bullet and the target create a critical mass that starts a chain reaction leading to a nuclear explosion.

A different design was used to detonate the bomb dropped on Nagasaki. Plutonium was stored in one large but subcritical mass. It was compressed to a critical density by means of surrounding chemical explosives. When the chemical explosive detonated, the blast forced the bomb material into a density that reached criticality. In either type of design, once criticality is reached the explosion follows in a millionth of a second.

In order for nuclear fission to occur, a bomb must use heavy atoms (isotopes) for fuel. Heavy atoms have many nucleons—neutrons and protons—in their nuclei. When these heavy nuclei split apart they release energy (and neutrons, which may cause nearby heavy nuclei to split apart also). Another more powerful type of nuclear weapon uses forms of hydrogen as fuel. Hydrogen has few subatomic particles in its nuclei—usually only a proton , but a proton plus a neutron in the isotope deuterium, and a proton plus two neutrons in the isotope tritium. Instead of splitting apart, these light atomic nuclei are forced together in high-speed collisions, a process called nuclear fusion. Energy is released when hydrogen nuclei fuse, forming helium. Fusion only occurs at temperatures of millions of degrees, such as exist in the hearts of stars. (The Sun and other stars generate their energy primarily by fusing hydrogen into helium.) On Earth only an atomic bomb can raise kilograms of material to such a temperature , which is why atomic bombs are used as detonators for hydrogen fusion bombs.

Because hydrogen is lighter than uranium, more hydrogen atoms fit into a sample of the same weight. Thus, even though one fusion reaction releases less energy than one fission reaction, more hydrogen than uranium atoms can be packed into a nuclear weapon and many more fusion reactions can take place in the weapon than fission reactions can take place in a fission bomb. Fusion weapons, therefore, produce bigger explosions than fission weapons of the same physical bulk.

By 1954, a new feature had been added to the hydrogen bomb to create an even more dangerous weapon. Like earlier hydrogen bombs, this weapon was detonated with the explosion of an atomic or fission weapon. This raised temperatures enough to cause the hydrogen atoms in the bomb to fuse and explode like a regular hydrogen bomb. The designers also enclosed this new bomb in a shell of uranium-238. Neutrons released from the fusion of hydrogen caused the uranium-238 in the surrounding jacket to undergo fission, adding to the power of the blast. This new device was, in effect, a fission-fusion-fission bomb.

The power or "yield" of a nuclear weapon is expressed in terms of how much TNT would be required to equal the weapon's blast. Units of kilotons (thousands of tons) and megatons (millions of tons) of TNT are used to describe nuclear blasts.


Effects of nuclear weapons

Nuclear weapons produce two important effects that are also produced by conventional, chemical explosives: they release heat and generate shock waves, pressure fronts of compressed air that smash objects in their paths. The heat released in a nuclear explosion creates a sphere of burning, glowing gas that can range from hundreds of feet to miles in diameter, depending on the power of the bomb. This fireball emits a flash of heat that travels outward from the site of the explosion (ground zero ), the area directly under the explosion. This heat can cause second degree burns to bare human flesh miles away from the blast site if the bomb is large enough. (Although this heat can start fires, it seems that much of the fire damage in Hiroshima and Nagasaki following the nuclear explosions resulted from damaged electrical, fuel, gas, and other systems following physical damage caused by the shock or blast wave that accompanied the explosion.)

The shock wave produced when a nuclear weapon explodes creates a front of moving air more powerful than any produced by a natural storm . Destructive winds follow the front of displaced air, causing more damage to objects in their path. Many nuclear weapons are designed to be detonated high above their targets to take advantage of this shock effect. The more powerful the bomb, the higher in the sky it will be detonated. The fission bombs dropped on Japan (Hiroshima, 13.5 kilotons; Nagasaki, 22 kilotons) exploded between 1,500 and 2,000 ft (458–610 m) above their targets. A bomb with the power of 10 megatons is capable of destroying most houses within a distance of more than 10 mi (16 km) from the blast site.

Unlike conventional explosives, nuclear devices can also release significant amounts of radioactivity and pulses of electromagnetic energy. Radioactivity is the release of fast particles and high-energy photons from unstable atomic nuclei. Besides the greater explosive power of nuclear weapons, radiation is the primary feature that most clearly distinguishes chemical from nuclear explosions. Radiation can kill outright at high doses and cause illnesses, including cancer , at lower doses. The initial burst of radiation during a nuclear explosion is made up of x rays , gamma rays, and neutrons. The energy of this radiation is so high that it can often penetrate buildings. Later, radioactive materials contaminate the explosion site and often enters the atmosphere where it can travel thousands of miles before falling back to earth. This source of radiation is called radioactive fallout . Radioactive fallout can harm living things for years following a nuclear explosion. Fission bombs and fission-fusion-fission bombs produce more fallout than hydrogen bombs because the fusion of hydrogen atoms generates less radioactive byproducts than does fission of uranium or plutonium.

Electromagnetic pulses (EMPs) are also produced by nuclear weapons that are exploded at high altitudes, and are caused by the interaction of radiation from the explosion with electrons in the atmosphere and with the Earth's magnetic field . EMPs are essentially powerful radio waves that can destroy many electronic circuits.

The effects of fires and destruction following a large-scale nuclear war might even change the climate of the planet . In 1983, a group of scientists including U.S. astronomer Carl Sagan (1934–1996) published the "nuclear winter" theory, which suggested that particles of smoke and dust produced by fires caused by many nuclear explosions would, for a time, block the Sun's rays from reaching the surface of Earth. This, in turn, would reduce temperatures and change wind patterns and ocean currents . These climatic changes, according to the theory, could destroy crops and lead to the death by famine of many more animals and humans than were killed outright by nuclear explosions. Some scientists have challenged these predictions, but others, including some United States government agencies, support them. On the other hand, there is no controversy about whether a large-scale nuclear war could kill hundreds of millions of people and imperil the future of modern civilization, even apart from nuclear winter effects.


Nuclear weapons today

Today nuclear weapons are built in many sizes and shapes not available in the 1940s and 1950s, and are designed for use against many different types of military and civilian targets. Some weapons are less powerful than 1,000 tons of TNT, while others have the explosive force of millions of tons of TNT. Small nuclear shells can be fired from cannons. Nuclear warheads mounted on missiles can be launched from land-based silos, ships, submarines, trains, and large wheeled vehicles. Several warheads can be fitted into one missile and directed to different targets in the same geographic area upon reentry into Earth's atmosphere. These multiple independently-targeted reentry vehicles (MIRVs) can release ten or so individual nuclear warheads far above their targets, making enemy interception more difficult and increasing the deadliness of each individual missile.

In general, nuclear weapons with "low" yields (in the kiloton, rather than the megaton, range) are termed "tactical," and are designed to be used in battle situations against specific military targets, such as a concentration of enemy troops or tanks, a naval vessel, or the like. These weapons are termed tactical because the word tactics, in military jargon, refers to the relatively small-scale maneuvers undertaken to win particular battles. Larger nuclear weapons are classed as "strategic," because the word strategy, again in military jargon, refers to the large-scale maneuvers undertaken to win whole wars. Strategic nuclear weapons are targeted mostly at cities and at other nuclear weapons, and are mostly designed to be dropped by bombers or launched on ballistic missiles ; tactical nuclear weapons are delivered by smaller devices over shorter distances. However, one nation's "tactical" warhead may be another's "strategic" warhead: Russia, for example, maintains that U.S. tactical warheads in Western Europe are in fact strategic warheads, since they can strike targets inside Russia itself, while Russian "tactical" warheads in the same arena cannot strike the U.S. heartland.

In the summer of 2002, President George W. Bush's administration sought and received permission from Congress to design a new class of nuclear weapons: "mini-nukes," relatively low-yield tactical nuclear weapons for use against underground bunkers and other small battlefield targets. Advocates of these new weapons point to the uniquely powerful, compact "punch" that can be delivered by a nuclear weapon; critics argue that even a small nuclear weapon may cause many civilian casualties, and, more important, that actual use of a nuclear weapon of any size would break the taboo on such use that has held since the end of World War II, making the use of larger, more destructive nuclear weapons more likely in future conflicts.

Even the ability of nuclear weapons to release radioactivity has been exploited to create different types of weapons. "Clean bombs" are weapons designed to produce as little radioactive fallout as possible. A hydrogen bomb without a uranium jacket would produce relatively little radioactive contamination , for example. A "dirty bomb" could just as easily be built, using materials that contribute to radioactive fallout. Such weapons could also be detonated near Earth's surface to increase the amount of material that could contribute to radioactive fallout. "Neutron" bombs originally designed to be used against Soviet forces in areas of Europe with cultural treasures (art museums, etc.) are able to shower battle fields with deadly neutrons that can penetrate buildings and armored vehicles without destroying them. Any people exposed to the neutrons, however, would die. (Neutron bombs also destroy with blast effects, but their deadly radiation zone extends far beyond their blast area)

The United States and Russia signed a Strategic Arms Reduction Treaty in 1993 to eliminate two thirds of their nuclear warheads in ten years. By 1995, nearly 2,500 nuclear warheads had been removed from bombers and missiles in the two countries, according to U.S. government officials. ("Elimination," in this context, does not necessarily mean dismantlement; many of the weapons that have been "eliminated" by treaty have been put in storage.) Although thousands of nuclear weapons still remain in the hands of many different governments, especially those of the U.S. and the Russian Federation, recent diplomatic trends have at least helped to lower the number of nuclear weapons in the world. This has caused many people to erroneously assume that the danger of nuclear weapons evaporated with the end of the Cold War.

However, the number of nations possessing nuclear weapons continues to increase, and the possibility of nuclear weapons being used against human beings for the first time since World War II may be larger than ever. In May 1995, more than 170 members of the United Nations agreed to permanently extend the Nuclear Non-Proliferation Treaty, first signed in 1960. Under terms of the treaty, the five major countries with nuclear weapons—the United States, Britain, France, Russia, and China—agreed to commit themselves to eliminating their arsenals as an "ultimate" goal. The other 165 signatory nations agree not to acquire nuclear weapons. Israel, which many Western intelligence assert possesses some nuclear weapons (but officially denies doing so), did not sign the treaty. Two other nuclear powers, India and Pakistan, also refused to renounce nuclear weapons. India and Pakistan, each of which probably possess several dozen nuclear weapons, have fought a number of border wars in recent decades, and in 2002 came frighteningly close, as many observers thought, to fighting a nuclear war. As of this writing (February 2003), North Korea has reactivated its nuclear-weapons-material production facilities and is engaged in a tense diplomatic standoff with the United States, which insists that North Korea abandon its nuclear-weapons program.

See also Atomic models; Nuclear power; Nuclear reactor.

Resources

books

Walmer, Max. An Illustrated Guide to Strategic Weapons. New York: Prentice Hall Press, 1988.


other

"U.S. Has 'Nuclear Hit List'." BBC News. March 2, 2002 [cited February 7, 2003] <http://news.bbc.co.uk/2/hi/americas/1864173.stm>.


K. Lee Lernerbr /> Larry Gilman
Dean Allen Haycock

KEY TERMS

. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Atomic bomb

—An explosive weapon which uses uranium–235 or plutonium as fuel. Its tremendous destructive power is produced by energy released from the "splitting of atoms" or nuclear fission. Also called A-bomb, atom bomb, or fission bomb.

Hydrogen bomb

—An nuclear explosive weapon which uses hydrogen isotopes as fuel and an atom bomb as a detonator. More powerful than an atom bomb, the Hydrogen bomb derives its destructive power from energy released when nuclei of hydrogen are forced together to form helium nuclei in a process called nuclear fusion. Also called H-bomb or Thermonuclear bomb.

Isotopes

—Two molecules in which the number of atoms and the types of atoms are identical, but their arrangement in space is different, resulting in different chemical and physical properties.

Nuclear fission

—"Splitting the atom." A nuclear reaction in which an atomic nucleus splits into fragments with the release of energy, including radioactivity.

Nuclear fusion

—A nuclear reaction in which an atomic nucleus combines with another nucleus and releases energy.

Nuclear weapon

—A bomb or other explosive that derives it explosive force from the release of nuclear energy. PlutoniumA heavy, rare natural element that undergoes fission in a nuclear bomb. It is produced artificially by bombarding uranium–238 with neutrons. The addition of one neutron to the nucleus of uranium–238 changes it into plutonium–239 which is called "weapons grade plutonium," the most efficient form for making weapons.

Radioactivity

—Spontaneous release of subatomic particles or gamma rays by unstable atoms as their nuclei decay.

Radioisotope

—A type of atom or isotope, such as strontium-90, that exhibits radioactivity.

TNT

—Trinitrotoluene, a high explosive.

Uranium

—A heavy natural element found in nature. More than 99% of natural uranium is a form called U–238. Only U–235 readily undergoes fission and it must be purified from the other form.

Nuclear Weapons

views updated May 21 2018

Nuclear Weapons. The possibility of creating nuclear weapons of almost unimaginable destructive power was first realized in the 1930s as physicists developed a fundamental understanding of the nucleus of the atom. A nuclear explosion is created when heavy nuclei are split—or fissioned—into several of their component parts that are smaller and more stable.

Impact of Nuclear Weapons.

Nuclear fission is a fundamentally different process from chemical explosions that occur in conventional high‐explosive or incendiary bombs. In chemical explosions, larger molecular structures are broken apart and rearranged into smaller parts, but the individual atomic nuclei remain untouched. A chemical explosion produces a sudden release of energy that generates an explosive blast, whose resulting high air pressures and strong winds can crush and knock down nearby structures and people. In the case of early nuclear weapons based on the fission process, the energy release, which occurs in microseconds, is enormously larger because the nuclear bonds that hold nuclei together and are broken during fission are so much stronger than the chemical bonds that bind atoms into molecules. Since the nuclear forces are typically 100,000 to 1 million times stronger than the electrical ones responsible for molecular structures, the resultant energy releases are correspondingly larger.

The nuclear blast is so powerful that it can crush objects many miles away with high winds in excess of 150 mph generated at distances greater than a mile. The release of the enormous energy in a nuclear explosion leads to extremely high temperatures, comparable to those that occur at the center of the Sun, causing massive and deadly fires. As a measure of comparison, the temperatures generated by nuclear weapons are hundreds to thousands of times higher than the temperatures on the surface of the Sun, which heats the surface of the Earth from a distance of more than 90 million miles. Dangerous radioactive fallout is also spread over large distances by the resulting nuclear radiation emerging with the nuclear debris.

The ability to release such enormous energy from single weapons, on a scale unparalleled in human history, profoundly alters the very nature of war, as well as its consequences. An appreciation of the consequences of a nuclear explosion can be learned from the experience of the only nuclear weapons used in war, the atomic bombs dropped by U.S. air forces on Hiroshima and Nagasaki in 1945. These two weapons devastated two entire cities. They had yields of 15–20 kilotons. That measure simply means that the energy release was the same as that from detonating 15,000–20,000 tons of TNT (TNT is an acronym for the chemical formula of dynamite). By way of comparison, the largest conventional bombs used in World War II—the so‐called blockbusters used by the Royal Air Force (RAF)—detonated 10 tons (20,000 pounds) of TNT.

Those fission bombs of 1945 are no more than primitive versions of the first stage, or triggers, of modern nuclear weapons, whose yields range into the megatons, or millions of tons of TNT equivalent, and whose deadly devastating impact ranges over many miles. (One kiloton is equivalent to 2 million pounds of TNT; 1 megaton is equivalent to 2 billion pounds of TNT.) In modern nuclear weapons, such fission triggers are known as the primaries. They ignite a secondary stage by creating very high temperatures in order to generate still larger quantities of energy by driving together, or fusing, light nuclei into more stable ones. This is known as fusion. Such modern weapons are commonly referred to as thermonuclear weapons—or, more simply, H‐bombs.

The effect of a 1‐megaton thermonuclear weapon has an energy release 100,000 times greater than the largest 10‐ton blockbusters of World War II; the area destroyed by blast would be several thousand times larger than that leveled by such blockbusters. Collateral destruction and casualties due to fires and radioactive fallout would extend even further than the area destroyed by blast.

Soon after World War II, it was realized that the existence of nuclear weapons posed a new and fearsome threat to modern civilization and that it was vital to treat them differently from “conventional”—nonnuclear—weapons. Serious initiatives during the decade immediately following WWII tried to bring these terrifying new weapons under international control. These efforts failed as the confrontation between the Western powers and the Soviet Union and its allies grew into a cold war. Fueled by this dangerous competition during the 1960s, the individual nuclear arsenals of the United States and the Soviet Union accumulated to tens of thousands of warheads. In addition, France, England, and China acquired their own, albeit much smaller, nuclear arsenals. Furthermore, the newly developed delivery systems of intercontinental‐range, and in particular, land‐based intercontinental ballistic missiles (ICBMs)—and long‐range ballistic missiles on submarines (SLBMs) moving about invisibly under the surface of the oceans—brought the threat of nuclear annihilation very close to home, less than thirty minutes away from a nation's borders.

Difficulty of Protection Against Nuclear Weapons.

It also became clear before long that there was no known or prospective technology that could provide a defense against a determined nuclear attack. In contrast to previous wars, essentially nothing would be left of a large urban “target”—its population and industry—if just one, or at most a few, nuclear warheads exploded over it. Witness the bombings of Hiroshima and Nagasaki.

A defense would have to be essentially perfect to provide protection against nuclear weapons, and that is neither a realistic standard of performance today nor a prospective one for future military systems. In World War II, during the Battle of Britain, the RAF defense system managed to destroy no more than one in ten of the attacking planes. At such a rate, the German Air Force was reduced faster than it could replace its losses. At the same time, cities like London could put out the fires and rebuild after the damage. Human defenselessness is a basic fact of the nuclear age. It is also troubling since it denies one of the most basic instincts of the human race: to defend ourselves, our families, our friends, our vital interests. Recognition of the ineffectiveness of defenses against the almost unimaginable destructive potential of a massive attack by nuclear bombs led the United States and the former Soviet Union to acknowledge that their very survival was based on mutual deterrence—ensuring that nuclear weapons were not used.

Basic Physical Processes in Nuclear Weapons.

The first step in detonating a thermonuclear weapon is to ignite the high explosive that causes a shock wave to travel inward and compress the nuclear material the explosive surrounds, known as the pit. At the same time, a strong source of neutrons is activated to flood the compressed pit.

If the material in the compressed pit reaches a condition known as criticality, the neutrons initiate a strong fission chain reaction. This is the fission, or primary, stage of a thermonuclear explosion. In a chain reaction, an incoming neutron splits the nucleus of fissile material (either an isotope of uranium, U235, that occurs in nature, or of plutonium, Pu239, that is man‐made), releasing at least two neutrons, which then run into other fissile material, producing more neutrons, which then run into other fissile material, and so on. Thus, in successive steps, or “generations,” of fission, the neutrons will multiply: 2, 2 × 2, 2 × 2 × 2, … After very roughly 100 generations, if the fissile material can be held together long enough, (i.e., for microseconds), enough nuclei will have fissioned and enough energy will have been created to generate an explosive equivalent to 10 kilotons or so of TNT.

Several years after the development of such first‐generation fission bombs, weapons designers concentrated on improving their performance by using the material more efficiently. U.S. and Soviet weapons technology advanced rapidly after the first Soviet nuclear detonation, “Joe 1,” in 1949. The biggest advance occurred when the process of fusion was introduced into the explosive process. Fusion, in contrast to fission, involves combining, or fusing together, several nuclei of the lightest elements, such as hydrogen isotopes, to form more stable heavy ones. High temperatures are required to ignite the fusion process effectively. This is because at high temperatures, individual nuclei acquire high speeds, and move sufficiently rapidly to push their way though their mutual electric repulsion and get near enough to each other to collide and “fuse” together. The new nucleus thus formed is generally more stable, leading to the release of a large energy, plus more neutrons. Fusion is the process fueling the Sun's burning.

Modern weapons with both fission and fusion stages are called thermonuclear or hydrogen bombs. In a thermonuclear weapon, the primary, or fission, stage creates the necessary high temperatures to ignite the fusion stage, which provides additional neutrons to initiate still more fission, thereby releasing much more energy. A thermonuclear weapon can be built with virtually no limit on the amount of fusion materials it contains. Such weapons generate explosions as large as tens of megatons of TNT, or the equivalent of billions of pounds of TNT. In thinking about the totality of destruction in a nuclear war waged with modern thermonuclear weapons of such enormous yield, it is well to keep in mind that many of the destructive effects of nuclear weapons were not anticipated, and were discovered with surprise by atomic scientists when they were used or tested. This calls for great humility when it comes to predicting the consequences of nuclear warfare.

Since 1945, the total number of known nuclear tests, worldwide, adds up to some 2,000. A major purpose of testing has been to validate and confirm appropriate performance specifications for new weapons types designed in response to military needs formulated during the Cold War. Starting in the mid‐1950s, U.S. weapons were designed and built “ready to go.” They conserved special nuclear materials (SNM)—the fissile materials Pu239 and U235—and were essentially maintenance‐free, ready to go at any time. “Ready” means that no physical changes or steps such as inserting the SNM had to be made in order to detonate a bomb. One merely had to launch and detonate the warhead by signal.

In response to growing worldwide concerns about radioactive fallout from continued nuclear testing, the United States, the Soviet Union, and the United Kingdom joined in 1963 in a Limited Test Ban Treaty that forbade testing aboveground, in the atmosphere, underwater, and in outer space. Only underground testing was allowed. A further restriction on testing was negotiated in 1974, limiting the yields of underground tests to a maximum of 150 kilotons, roughly ten times the yield of the Hiroshima bomb. This so‐called Threshold Test Ban Treaty was generally obeyed henceforth, though it was not ratified until 1990.

In 1992, progress in negotiated reductions in the nuclear arsenals, and further progress in reducing reliance on nuclear weapons after the end of the Cold War, led President George Bush to rule out nuclear weapons tests for new warheads and to declare a nine‐month moratorium on all nuclear testing. This moratorium was continued by his successor and has also been honored by Russia and the United Kingdom. On 11 August 1995, President Bill Clinton announced U.S. support for negotiating a comprehensive test ban treaty in 1996. The treaty would be of unending duration, and would include, as do all such tests, a “supreme national interest” clause should unanticipated circumstances present compelling arguments for renewed tests. Such arguments might arise if there were serious reversals from the present progress toward reducing nuclear danger in the world, or if unforeseen technical problems arose over time in the enduring nuclear stockpile.

By the best current technical judgment, U.S. weapons appear to be safe, reliable, age‐stable, and fully adequate for deterrence; but it will be a new challenge to maintain that confidence without being able to conduct tests that produce any nuclear yield. Under its recently formulated program for stockpile stewardship and management, the United States has accepted this challenge, following a comprehensive scientific review of prospects and needs for its nuclear arsenal. So have the United Kingdom, Russia, France, and China.

On September 1996 President Clinton was the first world leader to sign the Comprehensive Test Ban Treaty at the United Nations in New York. Soon thereafter the other declared nuclear powers—England, France, China, and Russia—also signed, and as of November 1998 150 nations have signed the Treaty and twenty‐one have ratified it. For it to go into effect it must be ratified by all forty‐four nuclear capable nations, i.e., nations with nuclear reactors for research or for civilian energy production, in addition to those with nuclear weapons. A Comprehensive Test Ban after more than 2,000 tests over a 50‐year period would be a tremendous achievement. Efforts to accomplish that goal are currently in progress, together with continuing efforts to reduce the size of the nuclear arsenals at the Strategic Arms Reduction Talks (START) underway between the U.S. and Russia.
[See also Arms Control and Disarmament: Nuclear; Cold War: External Course; Cold War: Domestic Course; War Plans; Weaponry; World War II: Military and Diplomatic Course.]

Bibliography

Margaret Gowing , Britain and Atomic Energy, 1939–1945, 1964.
Samuel Glasstone and Philip J. Dolan, eds., The Effects of Nuclear Weapons, 3rd ed. 1977.
Richard Rhodes , The Making of the Atomic Bomb, 1986.
Robert Serber , The Los Alamos Primer: The First Lectures on How to Build an Atomic Bomb, 1992.
David Holloway , Stalin and the Bomb, 1994.
Richard Rhodes , Dark Sun: The Making of the Hydrogen Bomb, 1995.

Sidney Drell

Nuclear Weapons

views updated Jun 11 2018

Nuclear weapons

Nuclear weapons are destructive devices that derive their power from nuclear reactions. The term weapon refers to devices such as bombs and warheads designed to deliver explosive power against an enemy. The two types of nuclear reactions used in nuclear weapons are nuclear fission and nuclear fusion. In nuclear fission, large nuclei are broken apart by neutrons, forming smaller nuclei, accompanied by the release of large amounts of energy. In nuclear fusion, small nuclei are combined with each other, again with the release of large amounts of energy.

Fission weapons

The design of a fission weapon is quite simple: all that is needed is an isotope that will undergo nuclear fission. Only three such isotopes exist: uranium-233, uranium-235, and plutonium-239. Fission occurs when the nuclei of any one of these isotopes is struck by a neutron. For example:

neutron + uranium-235 fission products + energy + more neutrons

The production of neutrons in this reaction means that fission can continue in other uranium-235 nuclei. A reaction of this kind is known as a chain reaction. All that is needed to keep a chain reaction going in uranium-235 is a block of the isotope of sufficient size. That size is called the critical size for uranium-235.

One of the technical problems in making a fission bomb is producing a block of uranium-235 (or other fissionable material) of exactly the right sizethe critical size. If the block is much less than the critical size, neutrons produced during fission escape to the surrounding air. Too few remain to keep a chain reaction going. If the block is larger than critical size, too many neutrons are retained. The chain reaction continues very rapidly and the block of uranium explodes before it can be dropped on an enemy.

The simplest possible design for a fission weapon, then, is to place two pieces of uranium-235 at opposite ends of a weapon casing. Springs are attached to each piece. When the weapon is delivered to the enemy (for example, by dropping a bomb from an airplane), a timing mechanism is triggered. At a given moment, the springs are released, pushing the two chunks of uranium-235 into each other. A piece of critical size is created, fission begins, and in less than a second the weapon explodes.

Words to Know

Fission bomb: An explosive weapon that uses uranium-235 or plutonium-239 as fuel. Also called an atom bomb.

Fusion bomb: An explosive weapon that uses hydrogen isotopes as fuel and an atom bomb as a detonator.

Isotopes: Two or more forms of an element that have the same chemical properties but that differ in mass because of differences in the number of neutrons in their nuclei.

Nuclear fission: A nuclear reaction in which an atomic nucleus splits into two or more fragments with the release of energy.

Nuclear fusion: A nuclear reaction in which two small atomic nuclei combine with each other to form a larger nucleus with the release of energy.

Radioactivity: The property possessed by some elements of spontaneously emitting energy in the form of particles or waves by disintegration of their atomic nuclei.

The only additional detail required is a source of neutrons. Even that factor is not strictly required since neutrons are normally present in the air. However, to be certain that enough neutrons are present to start the fission reaction, a neutron source is also included within the nuclear weapon casing.

Fusion weapons

A fusion weapon obtains the energy it releases from fusion reactions. Those reactions generally involve the combination of four hydrogen atoms to produce one helium atom. Such reactions occur only at very high temperatures, a few million degrees Celsius. The only way to produce temperatures of this magnitude on Earth is with a fission bomb. Thus, a fusion weapon is possible only if a fission bomb is used at its core.

Here is how the fusion bomb is designed: A fission bomb (like the one described in the preceding section) is placed at the middle of the fusion weapon casing. The fission bomb is then surrounded with hydrogen, often in the form of water, since water is two parts hydrogen (H2O). Even more hydrogen can be packed into the casing, however, if liquid hydrogen is used.

When the weapon is fired, the fission bomb is ignited first. It explodes, releasing huge amounts of energy and briefly raising the temperature inside the casing to a few million degrees Celsius. At this temperature, the hydrogen surrounding the fission bomb begins to fuse, releasing even larger amounts of energy.

The primary advantage that fusion weapons have over fission weapons is their size. Recall that the size of a fission explosion is limited by the critical size of the uranium-235 used in it. A weapon could conceivably consist of two pieces, each less than critical size; or three pieces, each less than critical size; or four pieces, each less than critical size, and so on. But the more pieces used in the weapon, the more difficult the design becomes. One must be certain that the pieces do not come into contact with each other and suddenly exceed critical size.

No such problem exists with a fusion bomb. Once the fission bomb is in place, the casing around it can be filled with ten pounds of hydrogen, 100 pounds of hydrogen, or 100 tons of hydrogen. The only limitation is how largeand heavythe designer wants the weapon to be.

The power difference between fission and fusion bombs is illustrated by the size of early models of each. The first fission bombs dropped on Japan at the end of World War II were rated as 20 kiloton bombs. The unit kiloton is used to rate the power of a nuclear weapon. It refers to the amount of explosive power produced by a thousand tons of the chemical explosive TNT. In other words, a 20-kiloton bomb has the explosive power of 20,000 tons of TNT. By comparison, the first fusion bomb ever tested had an explosive power of 5 megatons, or 5 million tons of TNT.

Effects of nuclear weapons

In some respects, the effects produced by nuclear weapons are similar to those produced by conventional chemical explosives. They release heat and generate shock waves. Shock waves are pressure fronts of compressed

air created as hot air expands away from the center of an explosion. They tend to crush objects in their paths. The heat released in a nuclear explosion creates a sphere of burning gas that can range from hundreds of feet to miles in diameter, depending on the power of the bomb. This fireball emits a flash of heat that travels outward from the site of the explosion or ground zero, the area directly under the explosion. The heat from a nuclear blast can set fires and cause serious burns to the flesh of humans and other animals.

Nuclear weapons also produce damage that is not experienced with chemical explosives. Much of the energy released during a weapons blast occurs in the form of X rays, gamma rays, and other forms of radiation that can cause serious harm to plant and animal life. In addition, the isotopes formed during fission and fusioncalled fission productsare all radioactive. These fission products are carried many miles away from ground zero and deposited on the ground, on buildings, on plant life, and on animals. As they decay over the weeks, months, and years following a nuclear explosion, the fission products continue to release radiation, causing damage to surrounding organisms.

Nuclear weapons today

Today nuclear weapons are built in many sizes and shapes. They are designed for use against various different types of military and civilian targets. Some weapons are rated at less than 1 kiloton in power, while others have the explosive force of millions of tons of TNT. Small nuclear shells can be fired from cannons. Nuclear warheads mounted on missiles can be launched from land-based silos, ships, submarine, trains, and large-wheeled vehicles. Several warheads can even be fitted into one missile. These MIRVs (or multiple independent reentry vehicles), can release up to a dozen individual nuclear warheads along with decoys far above their targets, making it difficult for the enemy to intercept them.

Even the ability of nuclear weapons to release radioactivity has been exploited to create different types of weapons. Clean bombs are weapons designed to produce as little radioactive fallout as possible. A hydrogen bomb without a uranium jacket would produce relatively little radioactive contamination, for example. A dirty bomb could just as easily be built with materials that contribute to radioactive fallout. Such weapons could also be detonated near Earth's surface to increase the amount of material that could contribute to radioactive fallout. Neutron bombs have been designed to shower battlefields with deadly neutrons that can penetrate buildings and armored vehicles without destroying them. Any people exposed to the neutrons, however, would die.

Radioactive Fallout

"The gift that keeps on giving."

That phrase is one way of describing radioactive fallout. Radioactive fallout is material produced by the explosion of a nuclear weapon or by a nuclear reactor accident. This material is blown into the atmosphere and then falls back to Earth over an extended period of time.

Radioactive fallout was an especially serious problem for about 20 years after the first atomic bombs were dropped in 1945. The United States and the former Soviet Union tested hundreds of nuclear weapons in the atmosphere. Each time one of these weapons was tested, huge amounts of radioactive materials were released to the atmosphere. They were then carried around the globe by the atmosphere's prevailing winds. Over long periods of time, they were carried back to Earth's surface or settled to the ground on their own (because of their weight).

More than 60 different kinds of radioactive materials are formed during the explosion of a typical nuclear weapon. Some of these decay and become harmless in a matter of minutes, hours, or days. Other remain radioactive for many years.

An example of a long-lived radioactive material is strontium-90. Strontium-90 loses one-half of its radioactivity every 28 years. It can continue to pose a threat, therefore, for more than a century. The special problem presented by strontium-90 is that it behaves very much like another elementcalcium. When it falls to Earth, it is taken up by grass, leaves, and other plant parts. When cows eat grass, they take in strontium-90. The strontium-90 is incorporated into their milk, which is then taken in by humans. Once in the human body, strontium-90 is incorporated into bones and teeth in much the same way that calcium is. Children growing up in the 1960s may still have low levels of strontium-90 in their systemsa "long-lasting gift" from the makers of nuclear weapons.

Nuclear weapons treaties

The United States and Russia signed a Strategic Arms Reduction Treaty (START I) in 1991, which called for the elimination of 9,000 nuclear warheads. Two years later, the two countries signed the START II Treaty, which called for the reduction of an additional 5,000 warheads beyond the number being reduced under START I. Under START II, each country agreed to reduce its total number of strategic nuclear warheads from bombers and missiles by two-thirds by 2003. In 1997, the United States and Russia agreed to delay the elimination deadline until 2007. By that time, each side must have reduced its number of nuclear warheads from 3,000 to 3,500.

Although thousands of nuclear weapons still remain in the hands of many different governments, recent diplomatic trends have helped to lower the number of nuclear weapons in the world. In May 1995, more than 170 members of the United Nations agreed to permanently extend the Nuclear Non-Proliferation Treaty (NPT), which was first signed in 1968. Under terms of the treaty, the five major countries with nuclear weaponsthe United States, Britain, France, Russia, and Chinaagreed to commit themselves to eliminating their arsenals as an ultimate goal and to refusing to give nuclear weapons or technology to any non-nuclear-weapon nation. The other 165 member nations agreed not to acquire nuclear weapons. Israel, which is believed to possess nuclear weapons, did not sign the treaty. Two other nuclear powers, India and Pakistan, refused to renounce nuclear weapons until they can be convinced their nations are safe without them. As of early 2000, a total number of 187 nations had agreed to the NPT. Cuba, India, Israel, and Pakistan were the only nations that had not yet agreed to the treaty.

[See also Nuclear fission; Nuclear fusion ]

Warfare, Nuclear

views updated Jun 08 2018

Warfare, Nuclear

COUNTERFORCE, COUNTERVALUE, AND MUTUAL ASSURED DESTRUCTION

MAINTAINING A STRATEGIC BALANCE

CURRENT CONCERNS AND FEARS

BIBLIOGRAPHY

Nuclear warfare consists of armed conflict between states in which one or more sides employ nuclear weapons. Because no war since World War II has involved nuclear weapons, how such a conflict would be triggered and executed is largely a matter of theoretical speculation. Furthermore, the sophistication and destructive scale of the nuclear weapons used against Japan pale in comparison to modern weapons. A nuclear war between two nuclear states would result in the deaths of hundreds of thousands, if not millions, of people. The areas surrounding locations hit with nuclear weapons would be highly contaminated with radioactive fallout. In addition, depending on the number of weapons used, such a war could have long-term devastating effects on the earths ecosystems and atmosphere. Because today only a few countries possess nuclear weapons, the number of conflicts that could conceivably escalate to nuclear war is limited. These countries include the United States, Russia, China, Great Britain, France, Israel, Pakistan, India, and most likely North Korea. Proliferation to additional countries remains a continual problem for international security.

The destructive power of nuclear weapons makes nuclear warfare fundamentally different from traditional conventional warfare. The single fifteen-kiloton bomb dropped on Hiroshima, for example, destroyed 80 percent of the city, immediately killing between 66,000 and 80,000 people and injuring roughly 70,000. As Wilfred Burchett (1945), the first journalist to report on the devastation, put it: Hiroshima does not look like a bombed city. It looks as if a monster steamroller has passed over it and squashed it out of existence. The city of Hiroshima estimates that the total killed from the explosion and subsequent radiation poisoning is over 240,000. Nagasaki saw high casualties as well, with 39,000 immediately killed and 25,000 injured, and many others who later died due to radiation poisoning.

How nuclear weapons would be used in war, and whether a nuclear war between two nuclear powers could even be won, has been a central problem facing military strategists and planners. Because of the devastating effects of nuclear weapons, they are less useful in battle than conventional weapons. However, because such weapons exist and because no country can be sure of what anothers intentions would be in the event that they were to gain a dominant nuclear advantage, the major nuclear powers have continued to develop nuclear war strategies. That said, nuclear powers have shown extreme caution when conflict develops with other nuclear powers, out of fear that a minor crisis could escalate into an unwanted nuclear war; this was displayed during the 1962 Cuban missile crisis. Nuclear powers have also been reluctant to use their nuclear capabilities in conflicts against a nonnuclear power, as with the United States in Vietnam or Israel in its 1973 war with Egypt and Syria.

COUNTERFORCE, COUNTERVALUE, AND MUTUAL ASSURED DESTRUCTION

Nuclear strategy makes distinctions between counterforce and countervalue. Counterforce strategies are intended to affect an opponents capabilities, whereas countervalue capabilities affect an opponents will. Counterforce targets an opponents armed forces and military-industrial installations, limiting the opponents ability to retaliate in a counterattack. A country that struck first in a nuclear war would most likely employ a counterforce targeting strategy. Countervalue strategies target an opponents citiesthat is, things of human and emotional value. A country that feared a nuclear attack by an opponent would threaten a countervalue retaliation with the hope that even the possibility of its opponent losing one city would be enough to deter a nuclear first strike. Of course, for a countervalue deterrent to be effective, the country being deterred must believe that at least some of its opponents nuclear arsenal would survive a first strike. It also must believe that the damage that that remaining arsenal could deliver would outweigh the benefits gained from striking first. With nuclear weapons it is oftentimes difficult to distinguish between what constitutes a counterforce and what constitutes a countervalue target. Military targets are often found in population centers and given the large radius of damage caused by a nuclear attack it is extremely difficult to target the one without hitting the other. For example, when U.S. war planners began looking for military-industrial targets across the Soviet Union after 1945, every sizeable Soviet city was deemed to contain military targets.

The logic behind counterforce and countervalue, as well as first-strike versus second-strike capabilities, is encompassed in the idea of mutual assured destruction (MAD). MAD describes a state of affairs in which both sides nuclear forces are such that a sufficient percent would remain after an attack that it would still be possible to bring about the near total destruction of the attacking state. The hope of MAD was that this mutual suicide pact would prevent either side from ever being tempted to use nuclear weapons. In order for MAD to be viable, however, it required the United States and the Soviet Union to stockpile large quantities of nuclear weapons and to develop targeting lists of single targets that would be hit multiple times. In addition, U.S. and Soviet force structures were designed to survive a possible first strike. Achieving this involved spending on difficult-to-target nuclear forces, such as submarines, hardened missile silos, and continually in-flight bomber fleets.

MAINTAINING A STRATEGIC BALANCE

Those who wanted to maintain a strategic nuclear balance put emphasis on developing less-accurate, single large warheads that would be unable to hit anything smaller than area targets. Such missiles would be effective against countervalue targets, which do not require precise accuracy to be effective; but would be less effective hitting silos or airfields.

It was feared, however, that a number of innovations and weapons systems could disrupt this strategic balance. Such a disruption could lead one side to perceive a window of opportunity in which they would be tempted to launch a preventive war before new technological innovations either restored the balance or shifted first-strike advantage to the opponent. For example, declarations of bomber gaps or missile gaps by United States politicians, particularly in the late 1950s and early 1960s, led many to fear that (alleged) Soviet advantage could lead to a devastating first strike. The development of multiple independently targeted reentry vehicles (MIRVs), which are intercontinental ballistic missiles (ICBMs) carrying multiple warheads that can be individually programmed to hit separate targets, was also seen as destabilizing, as one missile could target multiple ICBM silos. This offensive advantage, it was feared, could tempt one country to launch a preemptive attack out of fear that it would suffer a debilitating blow if it were not the one to attack first.

Another potential innovation capable of disrupting strategic balance is some form of missile defense system. While an effective missile defense system could protect a country from nuclear annihilation, it would also provide it with an overwhelming first-strike advantage, as its opponent would be unable to retaliate, regardless of the number of surviving nuclear forces. There would also be an incentive to strike sooner rather than later, as military history has shown that all defenses are eventually penetrable.

Arms control agreements between the United States and the Soviet Union during the Cold War were primarily designed to stabilize the strategic balance between the two sides. By limiting each sides ability to gain first-strike advantage, the hope was that neither side would be tempted to carry out a preemptive first strike. The Anti-Ballistic Missile (ABM) Treaty, Strategic Arms Limitation Talks (SALT I and II), the Strategic Arms Reduction Treaties (START I and II), and the Strategic Offensive Reduction Treaty (SORT) were all designed to provide a framework in which the United States and the Soviet Union (now Russia) could maintain a nuclear balance without engaging in a costly and potentially dangerous arms race.

CURRENT CONCERNS AND FEARS

Since the end of the Cold War, fears of a nuclear exchange between the United States and Russia have subsided. However, a number of concerns still remain. India and Pakistan, which both officially declared their nuclear status with a series of tests in 1997, have a long history of conflict, specifically over the contested Kashmir region. This history of conflict, their contingent border, and an underdeveloped command and control system, make a nuclear exchange (either intentional or accidental) a very real possibility.

It is also feared that a rogue state could develop a nuclear weapon and be able to hold the world hostage by threatening to use it against a major world city if its demands were not met. While the world could easily retaliate if such a threat were carried out, the question remains whether there would be a willingness to risk giving up an important city in the first place. It is this potentiality that has led world leaders to take aggressive stances (with mixed success) against such potential proliferators as North Korea, Iraq, Iran, and Libya.

The final fear is that a terrorist organization would be able to acquire a nuclear device by stealing, buying, or being given it from a countrys arsenal. This is a particularly difficult scenario because normal countervalue threats would not have a very strong deterrent effect on a small, decentralized, apocalyptic terrorist organization.

SEE ALSO Cold War; Defense; Defense, National; Deterrence, Mutual; Disarmament; Proliferation, Nuclear; World War II

BIBLIOGRAPHY

Allison, Graham T. 2004. Nuclear Terrorism: The Ultimate Preventable Catastrophe. New York: Times Books.

Burchett, Wilfred. 1945. The Atomic Plague. London Daily Express, September 4.

Larsen, Jeffrey A., and James M. Smith, eds. 2005. Historical Dictionary of Arms Control and Disarmament. Lanham, MD: Scarecrow Press.

Paul, T. V. 1995. Nuclear Taboo and War Initiation in Regional Conflicts. Journal of Conflict Resolution 39 (4): 696717.

Quester, George H. 2000. Nuclear Monopoly. New Brunswick, NJ: Transaction Publishers.

Schelling, Thomas C. 1966. Arms and Influence. New Haven, CT: Yale University Press.

Waltz, Kenneth N. 1990. Nuclear Myths and Political Realities. American Political Science Review 84 (3): 731745.

York, Herbert F. 1970. ABM, MIRV, and the Arms Race. Science 169 (3942): 257260.

David R. Andersen

Weaponry, Nuclear

views updated May 29 2018

Weaponry, Nuclear

THE EFFECTS OF A NUCLEAR EXPLOSION

DELIVERY METHODS

BIBLIOGRAPHY

The advent of the nuclear weapons age began on July 16, 1945, when the United States tested its first nuclear device in New Mexico at the Alamogordo Bombing Range, now known as the White Sands Missile Range. The successful nuclear explosion, named Trinity, was the end result of the Manhattan Project, a three-year, $1.9 billion ($26.9 billion in 2005 dollars) effort that brought hundreds of the worlds top scientists together to develop a weapon to be used in the United States war efforts against Japan and Germany. Nuclear weapons have been used in warfare on two occasions: on Hiroshima, Japan, on August 6, 1945, and on Nagasaki, Japan, on August 9, 1945. Both bombs were dropped by the United States. As of 2006, eight nations were known to possess nuclear weapons: the United States, Russia, the United Kingdom, France, China, Israel, India, and Pakistan. It is possible that North Korea also possesses a nuclear weapon. In 2003 North Korea claimed to have had successfully developed nuclear weapons. While North Korea has not tested a device, most intelligence estimates believe it is likely that it has nuclear capabilities. South Africa once possessed nuclear weapons but dismantled them in 1993 (see Cirincione, Wolfsthal, and Rajkumar 2005).

Nuclear weapons require fissionable materials. When a fissionable atom absorbs a neutron, it will split and release additional neutrons. In a nuclear chain reaction, those neutrons are absorbed into other fissionable atoms that subsequently split and release additional neutrons into other atoms. Nuclear explosions are the result of the rapid release of energy that comes from an uncontrolled nuclear chain reaction.

The two fissionable elements used in nuclear weapons are uranium and plutonium. Uranium is found in nature, but the specific fissionable isotope, uranium235, constitutes only 0.7 percent of all natural uranium. A nuclear weapon, however, requires uranium-235 to make up over 90 percent of the sample. In order to achieve such a high concentration, the uranium must go through an enrichment process that separates uranium-235 from the more common uranium-238 isotope. This has most commonly been achieved with centrifuges, but other methods, such as gaseous diffusion and electromagnetic isotope separation, have also been successful. Plutonium is not found in nature but is a product of the highly radioactive waste from a controlled chain reaction of uranium, usually performed in a nuclear reactor. To extract plutonium from this waste, a sophisticated chemical process is used. For a country seeking to establish a nuclear weapons program, these large-scale industrial and technical processes can be prohibitive.

Critical mass is the smallest amount of fissionable material that is needed to maintain a nuclear chain reaction. How much uranium or plutonium is needed to reach critical mass depends on various elements of weapon design, such as the shape of the fissile core (gun-type or sphere) or the effective use of reflectors to capture errant neutrons. Most estimates are that between 12 to 60 kilograms of weapons-grade uranium and 4 to 10 kilograms of plutonium are needed. In addition, the efficiency and yield of a weapon can be increased by adding a fusion fuel booster, such as lithium-6, as found in thermonuclear weapons.

THE EFFECTS OF A NUCLEAR EXPLOSION

The effects of a nuclear explosion are devastating. The majority of damage is caused by three main elements: blast effects, thermal heat, and ionizing radiation. For example, the bomb that was dropped on Hiroshima, a uranium-type device known as Little Boy, had a yield of 12.5 kilotons of TNT. Of the 76,000 buildings in Hiroshima, 48,000 were completely destroyed and another 22,000 were damaged. According to one study of the Hiroshima bombing, the temperature at the site of the explosion reached 5,400 degrees Fahrenheit and primary atomic bomb thermal injury was found in those exposed within [2 miles] of the hypocenter (quoted in Rhodes 1986, p. 714). The heat was so intense that people within a half mile of the fireball were reduced to bundles of smoking char. The number of deaths in Hiroshima due to the bomb is estimated to be 140,000, with an additional 60,000 dying from radiation effects over the next five years.

Since these early devices, the yield of nuclear weapons has grown considerably. Although never deployed, on October 30, 1961, the largest nuclear bomb ever tested was the Soviet Unions Tsar Bomba, which had a maximum yield of 100 megatons. More commonly, modern nuclear weapons have yields ranging between one and 5.5 megatons.

For a one-megaton device, the damage would be even more widespread than at Hiroshima and Nagasaki. According to Ansley J. Coale (1985), the shockwaves from a one-megaton blast would destroy modern multistory buildings within 2.9 miles and unreinforced brick and wood buildings within 4.2 miles of impact. Damage to brick and wood buildings would be substantial up to 8.5 miles from the blast. Heat would cause third-degree burns to exposed skin and set fire to clothing within 4.2 miles. The gamma rays produced from such a blast would be almost immediately lethal to any exposed person within 2.5 miles. People exposed at a slightly greater distance (2.7 miles) would have about a 50 percent mortality rate within a month of the explosion. Finally, a nuclear explosion that makes contact with the ground (as opposed to an airblast) would create tremendous amounts of radioactive fallout that could spread over an area as far as 1,000 square miles downwind from the explosion. Estimates of what percentage would be killed in a one-megaton blast on an urban population vary from 11 percent to 25 percent of the total population, with an additional 16 to 25 percent injured. Of course, in a nuclear exchange between advanced nuclear weapons states, multiple bombs would likely be assigned to single targets, resulting in even higher levels of devastation.

DELIVERY METHODS

The three main methods of delivery involve ballistic missiles, aircraft, and submarines. Delivery methods are tied to larger strategic and tactical issues related to nuclear deterrence. Nuclear states, such as the United States and the Soviet Union during the cold war, are concerned that a first-strike nuclear attack from another country could be so damaging that it would successfully eliminate any possibility for retaliation. As a result, states design their nuclear forces in such a way that a sufficient number of weapons would remain to respond with a devastating second strike. Many argue that the sole purpose of any nuclear weapon is to deter other states from ever using one. Some also fear that a terrorist organization could gain possession of a nuclear weapon and smuggle it into a major urban center.

Intercontinental ballistic missiles (ICBMs) are launched from reinforced below-ground silos and have ranges of more than 8,000 miles. Often, ICBMs are equipped with multiple warheadsmultiple, independently targeted reentry vehicles (MIRV)capable of hitting multiple targets. Shorter-range ballistic missiles, which could more easily be used in tactical or battlefield scenarios, have largely been eliminated from the arsenals of major nuclear states.

The appeal of aircraft and submarines is their mobility, as well as an enemys consequent difficulty in targeting them. Heavy-duty bombers, primarily equipped with up to twenty short-range attack missiles capable of hitting multiple targets, have the ability to penetrate enemy territory and withstand a great deal of abuse. Submarines carrying strategic nuclear missiles can remain below the surface for long periods and can launch missiles capable of hitting specific targets over distances of hundreds of miles. The possession of a nuclear-equipped submarine fleet gives a country a very credible second-strike deterrent.

Since the end of the cold war, both the United States and the former Soviet Union have worked to decrease their nuclear arsenals. However, many fear that tensions between other nuclear states, such as India and Pakistan, and the ongoing threat of further proliferation could result in the future use of nuclear weapons.

SEE ALSO Defense; Defense, National; Deterrence, Mutual; Disarmament; Proliferation, Nuclear; World War II

BIBLIOGRAPHY

Barnaby, Frank. 2003. How to Build a Nuclear Bomb and Other Weapons of Mass Destruction. London: Granta.

Campbell, Christopher. 1984. Nuclear Weapons Fact Book. Novato, CA: Presidio Press.

Cirincione, Joseph, with Jon B. Wolfsthal and Miriam Rajkumar. 2005. Deadly Arsenals: Tracking Weapons of Mass Destruction. Washington, DC: Carnegie Endowment for International Peace.

Coale, Ansley J. 1985. Nuclear War and Demographers Projections. Population and Development Review 11 (3): 483493.

Nuclear Weapons Data. Bulletin of the Atomic Scientists. http://www.thebulletin.org/nuclear_weapons_data.

Rhodes, Richard. 1986. The Making of the Atomic Bomb. New York: Simon and Schuster.

Schwartz, Stephen I., ed. 1998. Atomic Audit: The Costs and Consequences of U.S. Nuclear Weapons Since 1940. Washington, DC: Brookings Institution Press.

David R. Andersen

About this article

nuclear weapons

All Sources -
Updated Aug 13 2018 About encyclopedia.com content Print Topic