You are on page 1of 13

Simplified representation

Many particles appear in reactions so often that they are usually abbreviated. Thus, a
helium nucleus (also known as an alpha particle) is written with the Greek letter "α".
Deuterons (heavy hydrogen, 21H) are written simply as "d". Also, because the atomic
numbers are implied by the chemical symbols, they are redundant after balancing, and
often omitted. Finally, many common reactions take the form of a relatively heavy
nucleus being struck by one of a small group of common, reactive particles, and
emitting another common particle, to produce another nucleus. For these reactions,
the notation can be greatly condensed into the form:
<target nucleus> ( incoming particle, outgoing particle(s) ) <product nucleus>
So we could rewrite the preceding example by first abbreviating symbols:
63Li + d → α + α
then dropping atomic numbers:
6Li + d → α + α
and finally using the condensed form:

6Li(d,α)α
Reaction rates
The rate at which reactions occur depends on the particle flux and the reaction cross
section. See those articles for more information.

Neutrons versus ions


In the initial collision which begins the reaction, the particles must approach closely
enough so that the short range strong force can affect them. As most common nuclear
particles are positively charged, this means they must overcome considerable
electrostatic repulsion before the reaction can begin. Even if the target nucleus is part
of a neutral atom, the other particle must penetrate well beyond the electron cloud and
closely approach the nucleus, which is positively charged. Thus, such particles must
be first accelerated to high energy, for example by:
 particle accelerators
 nuclear decay (alpha particles are the main type of interest here, since beta and
gamma rays are rarely involved in nuclear reactions)
 very high temperatures, on the order of millions of degrees, producing
thermonuclear reactions
 cosmic rays

Also, since the force of repulsion is proportional to the product of the two charges,
reactions between heavy nuclei are rarer, and require higher initiating energy, than
those between a heavy and light nucleus; while reactions between two light nuclei are
commoner still.
Neutrons, on the other hand, have no electric charge to cause repulsion, and are able
to effect a nuclear reaction at very low energies. In fact at extremely low particle
energies (corresponding, say, to thermal equilibrium at room temperature), the
neutron's de Broglie wavelength is greatly increased, possibly greatly increasing its
capture cross section. Thus low energy neutrons may be even more reactive than high
energy neutrons.

Notable types
While the number of possible nuclear reactions is immense, there are several types
which are more common, or otherwise notable. Some examples include:
 Fusion reactions - two light nuclei join to form a heavier one, with additional
particles (usually protons or neutrons) thrown off to conserve momentum.
 Fission reactions - a very heavy nucleus absorbs additional light particles
(usually neutrons) and splits into two roughly equal-sized pieces.
 Spallation - a nucleus is hit by a very high energy particle, and is smashed into
many fragments.
 (d,n) and (d,p) reactions. A deuteron beam impinges on a target; the target
nuclei absorb either the neutron or proton from the deuteron.
 (α,n) and (α,p) reactions. Some of the earliest nuclear reactions studied
involved an alpha particle produced by alpha decay, knocking a nucleon from
a target nucleus.

History
The process was discovered in 1939 by Otto Hahn, Lise Meitner and coworkers at the
Kaiser-Wilhelm-Institute for Chemistry in Berlin.
The results of the bombardment of uranium by neutrons had proved interesting and
puzzling. First studied by Fermi and his colleagues in 1934, they were not properly
interpreted until several years later.
On January 16 1939, Niels Bohr of Copenhagen, Denmark, arrived in the United
States to spend several months in Princeton, N. J., and was particularly anxious to
discuss some abstract problems with Albert Einstein. (Four years later Bohr was to
escape to Sweden from Nazi-occupied Denmark in a small boat, along with thousands
of other Danish Jews, in large scale operation.) Just before Bohr left Denmark, two of
his colleagues, Otto Robert Frisch and Lise Meitner (both refugees from Germany),
had told him their guess that the absorption of a neutron by a uranium nucleus
sometimes caused that nucleus to split into approximately equal parts with the release
of enormous quantities of energy, a process that they dubbed nuclear "fission."
The occasion for this hypothesis was the important discovery of Otto Hahn and Fritz
Strassmann in Germany (published in Naturwissenschaften in early January 1939)
which proved that an isotope of barium was produced by neutron bombardment of
uranium. Bohr had promised to keep the Meitner/Frisch interpretation secret until
their paper was published to preserve priority, but on the boat he discussed it with
Leon Rosenfeld, but forgot to tell him to keep it secret. Rosenfeld immediately upon
arrival told everyone at Princeton University, and from them the news spread by word
of mouth to neighboring physicists including Enrico Fermi at Columbia University.
As a result of conversations among Fermi, John R. Dunning, and G. B. Pegram, a
search was undertaken at Columbia for the heavy pulses of ionization that would be
expected from the flying fragments of the uranium nucleus. On January 26, 1939,
there was a conference on theoretical physics at Washington, D. C., sponsored jointly
by the George Washington University and the Carnegie Institution of Washington.
Fermi left New York to attend this meeting before the Columbia fission experiments
had been tried. At the meeting Bohr and Fermi discussed the problem of fission, and
in particular Fermi mentioned the possibility that neutrons might be emitted during
the process. Although this was only a guess, its implication of the possibility of a
chain reaction was obvious. A number of sensational articles were published in the
press on this subject. Before the meeting in Washington was over, several other
experiments to confirm fission had been initiated, and positive experimental
confirmation was reported from four laboratories (Columbia University, Carnegie
Institution of Washington, Johns Hopkins University, University of California) in the
February 15 1939, issue of the Physical Review. By this time Bohr had heard that
similar experiments had been made in his laboratory in Copenhagen about January 15.
(Letter by Frisch to Nature dated January 16 1939, and appearing in the February 18
issue.) Frédéric Joliot in Paris had also published his first results in the Comptes
Rendus of January 30 1939. From this time on there was a steady flow of papers on
the subject of fission, so that by the time (December 6 1939) L. A. Turner of
Princeton wrote a review article on the subject in the Reviews of Modern Physics
nearly one hundred papers had appeared. Complete analysis and discussion of these
papers have appeared in Turner's article and elsewhere.

Effects of isotopes
Natural uranium contains three isotopes: U-234 (0.006%), U-235 (0.7%), and U-238
(99.3%). The speed required for a fission event vs. non-fission capture event is
different for different isotopes.
U-238 tends to capture intermediate speed neutrons (creating U-239, not fission).
High speed neutrons tend to have inelastic collisions with U-238, which just slow
down the neutrons. Thus, U-238 tends both to reduce the speed of the fast neutrons
and then capture them when they get to an intermediate speed.
U-235 fissions with a much wider range of neutron speeds than U-238. Since U-238
affects many neutrons without inducing fission, having it in the mix is bad for
promoting fission. So, if we separate the U-235 from the U-238 and discard the U-238
(producing enriched uranium), we promote a chain reaction. In fact, the probability of
fission of U-235 by high speed neutrons may be great enough to make the use of a
moderator unnecessary once the U-238 has been removed.
U-235 is present in natural uranium only to the extent of about one part in 140. Also,
the relatively small difference in mass between the two isotopes makes isotope
separation difficult. Nevertheless, the possibility of separating U-235 was recognized
early on in the Manhattan Project as being of the greatest importance to their success.

Reduction of non-fission capture by isotope


separation
An additional complication is that natural uranium contains three isotopes: U-234, U-
235, and U-238, present to the extent of approximately 0.006, 0.7, and 99.3 per cent,
respectively. We have already seen that the probabilities of processes (2) and (4) are
different for different isotopes. We have also seen that the probabilities are different
for neutrons of different energies.
For neutrons of certain intermediate speeds (corresponding to energies of a few
electron volts) U-238 has a large capture cross section for the production of U-239 but
not for fission. There is also a considerable probability of inelastic (i.e., non-capture-
producing) collisions between high speed neutrons and U-238 nuclei. Thus the
presence of the U-238 tends both to reduce the speed of the fast neutrons and to effect
the capture of those of moderate speed. Although there may be some non-fission
capture by U-235, it is evident that if we can separate the U-235 from the U-238 and
discard the U-238, we can reduce non-fission capture and can thus promote the chain
reaction. In fact, the probability of fission of U-235 by high speed neutrons may be
great enough to make the use of a moderator unnecessary once the U-238 has been
removed. Unfortunately, U-235 is present in natural uranium only to the extent of
about one part in 140. Also, the relatively small difference in mass between the two
isotopes makes separation difficult. Nevertheless, the possibility of separating U-235
was recognized early on in the Manhattan Project as being of the greatest importance.

Production and purification of materials


It has been stated above that the cross section for capture of neutrons varies greatly
among different materials. In some it is very high compared to the maximum fission
cross section of uranium. If, then, we are to hope to achieve a chain reaction, we must
reduce effect (3) - non-fission capture by impurities -to the point where it is not
serious. This means very careful purification of the uranium metal and very careful
purification of the moderator. Calculations show that the maximum permissible
concentrations of many impurity elements are a few parts per million- in either the
uranium or the moderator. When it is mentioned that up to 1940 the total amount of
uranium metal produced in the USA was not more than a few grams and even this was
of doubtful purity, that the total amount of metallic beryllium produced in the USA
was not more than a few kilograms, that the total amount of concentrated deuterium
produced was not more than a few kilograms, and that carbon had never been
produced in quantity with anything like the purity required of a moderator, it is clear
that the problem of producing and purifying materials was a major one.
The problem of producing large amounts of high purity uranium was solved by Frank
Spedding using the thermite process. Ames Laboratory was established in 1942 to
produce the large amounts of uranium that would be necessary for the research to
come.

Neutronicity, confinement requirement, and power density

Any of the reactions above can in principle be the basis of fusion power production.
In addition to the temperature and cross section discussed above, we must consider
the total energy of the fusion products Efus, the energy of the charged fusion products
Ech, and the atomic number Z of the non-hydrogenic reactant.
Specification of the D-D reaction entails some difficulties, though. To begin with, one
must average over the two branches (2) and (3). More difficult is to decide how to
treat the T and 3He products. T burns so well in a deuterium plasma that you probably
can't get it out even if you want to. The D-3He reaction is optimized at a much higher
temperature, so the burnup at the optimum D-D temperature may be low, so it seems
reasonable to assume the T but not the 3He gets burned up and adds its energy to the
net reaction. Thus we will count the DD fusion energy as Efus = (4.03+17.6+3.27)/2
= 12.5 MeV and the energy in charged particles as Ech = (4.03+3.5+0.82)/2 = 4.2
MeV.
Another unique aspect of the D-D reaction is that there is only one reactant, which
must be taken into account when calculating the reaction rate.
With this choice, we tabulate parameters for four of the most important reactions.

fuel Z Efus [MeV] Ech [MeV] neutronicity

D-T 1 17.6 3.5 0.80

D-D 1 12.5 4.2 0.66

D-3He 2 18.3 18.3 ~0.05

p-11B 5 8.7 8.7 ~0.001

The last column is the neutronicity of the reaction, the fraction of the fusion energy
released as neutrons. This is an important indicator of the magnitude of the problems
associated with neutrons like radiation damage, biological shielding, remote handling,
and safety. For the first two reactions it is calculated as (Efus-Ech)/Efus. For the last
two reactions, where this calculation would give zero, the values quoted are rough
estimates based on side reactions that produce neutrons in a plasma in thermal
equilibrium.
Of course the reactants should also be mixed in the optimal proportions. This is the
case when each reactant ion plus its associated electrons accounts for half the
pressure. Assuming that the total pressure is fixed, this means that density of the non-
hydrogenic ion is smaller than that of the hydrogenic ion by a factor 2/(Z+1).
Therefore the rate for these reactions is reduced by the same factor, on top of any
differences in the values of <σv>/T². On the other hand, because the D-D reaction has
only one reactant, the rate is twice as high as if the fuel were divided between two
hydrogenic species.
Thus there is a "penalty" of (2/(Z+1)) for non-hydrogenic fuels arising from the fact
that they require more electrons, which take up pressure without participating in the
fusion reaction. There is at the same time a "bonus" of a factor 2 for D-D due to the
fact that each ion can react with any of the other ions, not just a fraction of them.
We can now compare these reactions in the following table.

fuel <σv>/T² penalty/bonus reactivity Lawson criterion power density

D-T 1.24e-24 1 1 1 1

D-D 1.28e-26 2 48 30 68

D-3He 2.24e-26 2/3 83 16 80

p-11B 3.01e-27 1/3 1240 500 2500

The maximum value of <σv>/T² is taken from a previous table. The "penalty/bonus"
factor is that related to a non-hydrogenic reactant or a single-species reaction. The
values in the column "reactivity" are found by dividing (1.24e-24) by the product of
the second and third columns. It indicates the factor by which the other reactions
occur more slowly than the D-T reaction under comparable conditions. The column
"Lawson criterion" weights these results with Ech and gives an indication of how
much more difficult it is to achieve ignition with these reactions, relative to the
difficulty for the D-T reaction. The last column is label "power density" and weights
the practical reactivity with Efus. It indicates how much lower the fusion power
density of the other reactions is compared to the D-T reaction and can be considered a
measure of the economic potential.

Bremsstrahlung losses

The ions undergoing fusion will essentially never occur alone but will be mixed with
electrons that neutralize the ions' electrical charge and form a plasma. The electrons
will generally have a temperature comparable to or greater than that of the ions, so
they will collide with the ions and emit Bremsstrahlung. The Sun and stars are opaque
to Bremsstrahlung, but essentially any terrestrial fusion reactor will be optically thin
at relevant wavelengths. Bremsstrahlung is also difficult to reflect and difficult to
convert directly to electricity, so the ratio of fusion power produced to
Bremsstrahlung radiation lost is an important figure of merit. This ratio is generally
maximized at a much higher temperature than that which maximizes the power
density (see the previous subsection). The following table shows the rough optimum
temperature and the power ratio at that temperature for several reactions.
[2] (http://theses.mit.edu/Dienst/UI/2.0/Page/0018.mit.theses/1995-130/26?
npages=306)

fuel Ti (keV) Pfusion/PBremsstrahlung

D-T 50 140

D-D 500 2.9

D-3He 100 5.3

3He-3He 1000 0.72

p-6Li 800 0.21

p-11B 300 0.57

The actual ratios of fusion to Bremsstrahlung power will likely be significantly lower
for several reasons. For one, the calculation assumes that the energy of the fusion
products is transmitted completely to the fuel ions, which then lose energy to the
electrons by collisions, which in turn lose energy by Bremsstrahlung. However
because the fusion products move much faster than the fuel ions, they will give up a
significant fraction of their energy directly to the electrons. Secondly, the plasma is
assumed to be composed purely of fuel ions. In practice, there will be a significant
proportion of impurity ions, which will lower the ratio. In particular, the fusion
products themselves must remain in the plasma until they have given up their energy,
and will remain some time after that in any proposed confinement scheme. Finally, all
channels of energy loss other than Bremsstrahlung have been neglected. The last two
factors are related. On theoretical and experimental grounds, particle and energy
confinement seem to be closely related. In a confinement scheme that does a good job
of retaining energy, fusion products will build up. If the fusion products are efficiently
ejected, then energy confinement will be poor, too.
The temperatures maximizing the fusion power compared to the Bremsstrahlung are
in every case higher than the temperature that maximizes the power density and
minimizes the required value of the fusion triple product. This will not change the
optimum operating point for D-T very much because the Bremsstrahlung fraction is
low, but it will push the other fuels into regimes where the power density relative to
D-T is even lower and the required confinement even more difficult to achieve. For
D-D and D-3He, Bremsstrahlung losses will be a serious, possibly prohibitive
problem. For 3He-3He, p-6Li and p-11B the Bremsstrahlung losses appear to make a
fusion reactor using these fuels impossible.
Critical mass in nuclear fission (left over)

While any critical mass will in principle lead to exponential growth, the time this will
take depends on several factors. The degree to which the mass is supercritical affects
the rate of growth. However, as mentioned above, a fraction of the neutrons that cause
fission do so only after a brief delay. This delay slows the process of exponential
growth and permits the control of nuclear chain reactions. If there are enough
neutrons captured so that the ones causing immediate fission are sufficient to lead to
exponential growth, then the mass is called prompt critical and it becomes very
difficult to control.
A simple nuclear weapon relies on this exponential growth to induce fission in a
significant fraction of the fissile nuclei it contains. Such a device must not only be
prompt critical, it must be highly prompt critical. Moreover, it must be rapidly
converted from a subcritical configuration (for storage) to a highly prompt critical
configuration upon detonation. This is a difficult procedure; see nuclear weapon
design for an overview.
The relative number of neutrons which escape from a quantity of uranium can be
minimized by changing the size and shape. In a sphere any surface effect is
proportional to the square of the radius, and any volume effect is proportional to the
cube of the radius. Now the escape of neutrons from a quantity of uranium is a surface
effect depending on the area of the surface, but fission capture occurs throughout the
material and is therefore a volume effect. Consequently the greater the amount of
uranium, the less probable it is that neutron escape will predominate over fission
capture and prevent a chain reaction. Loss of neutrons by non-fission capture is a
volume effect like neutron production by fission capture, so that increase in size
makes no change in its relative importance..

Moderators in nuclear fission (left over)

It occurred to a number of physicists that it might be possible to mix uranium with a


moderator in such a way that the high speed fission neutrons, after being ejected from
uranium and before re-encountering uranium nuclei, would have their speeds reduced
below the speeds for which non-fission capture is highly probable. The characteristics
of a good moderator are that it should be of low atomic weight and that it should have
little or no tendency to absorb neutrons. Lithium and boron are excluded on the latter
count. Helium is difficult to use because it is a gas and forms no compounds. The
choice of moderator therefore lay (and still may lie) among hydrogen, deuterium,
beryllium, and carbon. It was Enrico Fermi and Leó Szilárd who first proposed the
use of graphite (a form of carbon) as a moderator for a chain reaction.

Control - weapons or power?


The problems that have been discussed so far have to do merely with the realization
of the chain reaction. If such a reaction is going to be of use, we must be able to
control it. The problem of control is different depending on whether we are interested
in steady production of power or in an explosion. In general, the steady production of
atomic power requires a slow-neutron-induced fission chain reaction occurring in a
mixture or lattice of uranium and moderator, while an atomic bomb requires a fast-
neutron-induced fission chain reaction in U-235 or Pu-239, although both slow- and
fast-neutron fission may contribute in each case. It seemed likely even in 1940, that
by using neutron absorbers a power chain reaction could be controlled. It was also
considered likely, though not certain, that such a chain reaction would be self-limiting
by virtue of the lower probability of fission-producing capture when a higher
temperature was reached. Nevertheless, there was a possibility that a chain-reacting
system might get out of control, and it therefore seemed necessary to perform the
chain-reaction experiment in an uninhabited location.
Up to this point we have been discussing how to produce and control a nuclear chain
reaction but not how to make use of it. The technological gap between producing a
controlled chain reaction and using it as a large-scale power source or an explosive is
comparable to the gap between the discovery of fire and the manufacture of a steam
locomotive.
Although production of power has never been the principal object of this project,
enough attention has been given to the matter to reveal the major difficulty: the
attainment of high-temperature operation. An effective heat engine must not only
develop heat but must develop heat at a high temperature. To run a chain-reacting
system at a high temperature and to convert the heat generated to useful work is very
much more difficult than to run a chain-reacting system at a low temperature.
Of course, the proof that a chain reaction is possible does not itself ensure that nuclear
energy can be effective in a bomb. To have an effective explosion it is necessary that
the chain reaction build up extremely rapidly; otherwise only a small amount of the
nuclear energy will be utilized before the bomb flies apart and the reaction stops. It is
also necessary that no premature explosion occur. This entire "detonation" problem
was and still remains one of the most difficult problems in designing a high-efficiency
atomic bomb.
Three ways of increasing the likelihood of a chain reaction have been mentioned: use
of a moderator; attainment of high purity of materials; and use of special material,
either U-235 or Pu-239. The three procedures are not mutually exclusive, and many
schemes have been proposed for using small amounts of separated U-235 or Pu-239
in a lattice composed primarily of ordinary uranium or uranium oxide and of a
moderator or two different moderators. Such proposed arrangements are usually
called "enriched piles".

Nuclear fusion

It takes considerable energy to force nuclei to fuse, even those of the least massive
element, hydrogen. But the fusion of lighter nuclei, which creates a heavier nucleus
and a free neutron, will generally release even more energy than it took to force them
together — an exothermic process that can produce self-sustaining reactions.

The energy released in most nuclear reactions is much larger than that for chemical
reactions, because the binding energy that glues a nucleus together is far greater than
the energy that holds electrons to a nucleus. For example, the ionization energy
gained by adding an electron to hydrogen is 13.6 electron volts -- less than one-
millionth of the 17 MeV released in the D-T reaction shown below.

Methods of fuel confinement

The fusion reaction can sustain itself if enough of the energy produced goes into
keeping the fuel hot.
Gravitational confinement One force capable of confining the fuel well enough to
satisfy the Lawson criterion is gravity. The mass needed, however, is so great that
gravitational confinement is only found in stars. Even if the more reactive fuel
deuterium were used, a mass about the size of the Moon would be needed.
Magnetic confinement Since plasmas are very good electrical conductors, magnetic
fields can also be used to confine fusion fuel. A variety of magnetic configurations
can be used, the most basic distinction being between mirror confinement and toroidal
confinement, especially tokamaks and stellarators.
Inertial confinement A third confinement principle is to apply a rapid pulse of energy
to a measure of fusion fuel, causing it to simultaneously "implode" and heat to very
high pressure and temperature. If the fuel is dense enough and hot enough, the fusion
reaction rate will be high enough to burn a significant fraction of the fuel before it has
dissipated. To achieve these extreme conditions, the initially cold fuel must be
explosively compressed. Inertial confinement is used in the hydrogen bomb, where
the driver is x-rays created by a fission bomb, but is also attempted in "controlled"
nuclear fusion, where the driver is a laser, ion, or electron beam.
Some other confinement principles have been investigated, such as muon-catalyzed
fusion, the Farnsworth-Hirsch fusor (inertial electrostatic confinement), and bubble
fusion.

There are two effects that lower the actual temperature needed. One is the fact that
temperature is the average kinetic energy, implying that some nuclei at this
temperature would actually have much higher energy than 0.1 MeV, while others
would be much lower. It is the nuclei in the high-energy tail of the velocity
distribution that account for most of the fusion reactions. The other effect is quantum
tunneling. The nuclei do not actually have to have enough energy to overcome the
Coulomb barrier completely. If they have nearly enough energy, they can tunnel
through the remaining barrier. For this reason fuel at lower temperatures will still
undergo fusion events, at a lower rate.

The reaction cross section σ is a measure of the probability of a fusion reaction as a


function of the relative velocity of the two reactant nuclei. If the reactants have a
distribution of velocities, e.g. a thermal distribution with thermonuclear fusion, then it
is useful to perform an average of over the distributions of the product of cross section
and velocity. The reaction rate (fusions per volume per time) is <σv> times the
product of the reactant number densities:

If a species of nuclei is reacting with itself, such as the DD reaction, then the product
n1n2 must be replaced by (1 / 2)n2.
increases from virtually zero at room temperatures up to meaningful
magnitudes at temperatures of 10 - 100 keV. At these temperatures, well above
typical ionization energies (13 eV in the hydrogen case), the fusion reactants exist in a
plasma state.
The significance of <σv> as a function of temperature in a device with a particular
energy confinement time is found by considering the Lawson criterion

Important fusion reactions


Astrophysical reaction chains

The most important fusion process in nature is that which powers the stars. The net
result is the fusion of four protons into one alpha particle, with the release of two
electrons, two neutrinos, and energy, but several individual reactions are involved,
depending on the mass of the star. For stars the size of the sun or smaller, the proton-
proton chain dominates. In heavier stars, the CNO cycle is more important.

Criteria and candidates for terrestrial reactions

In man-made fusion, the primary fuel is not constrained to be protons and higher
temperatures can be used, so reactions with larger cross-sections are chosen. This
implies a lower Lawson criterion, and therefore less startup effort. Another concern is
the production of neutrons, which activate the reactor structure radiologically, but also
have the advantages of allowing volumetric extraction of the fusion energy and
tritium breeding. Reactions that release no neutrons are referred to as aneutronic.
In order to be useful as a source of energy, a fusion reaction must satisfy several
criteria. It must:
 ... be exothermic. This one is obvious, but it limits the reactants to the low Z
side of the curve of binding energy. It also makes helium He-4 the most
common product because of its extraordinarily tight binding, although He3
and T also show up.
 ... involve low Z nuclei. This is because the electrostatic repulsion must be
overcome before the nuclei are close enough to fuse.
 ... have two reactants. At anything less than stellar densities, three body
collisions are too improbable.
 ... have two or more products. This allows simultaneous conservation of
energy and momentum without relying on the (weak!) electromagnetic force.
 ... and conserve both protons and neutrons. The cross sections for the weak
interaction are too small.
Not many reactions meet these criteria. The most interesting reactions are the
following.

(3.5 (14.1
(1) D +T →   4He +  n  
MeV) MeV)
(1.01 (3.02
(2) D +D →  T +  p         (50%)
MeV) MeV)
(0.82 (2.45
(3)       →   3He +  n         (50%)
MeV) MeV)
(3.6 (14.7
(4) D + 3He →   4He +  p
MeV) MeV)
+ 11.3
(5) T +T →   4He   + 2  n
MeV
+ 12.9
(6) 3He + 3He →   4He   + 2  p
MeV
+ 12.1
(7) 3He + T →   4He   +  p   +n   (51%)
MeV
(4.8 (9.5
(8)       →   4He +  D         (43%)
MeV) MeV)
(0.5 (1.9 (11.9
(9)       →   4He +  n +p   (6%)
MeV) MeV) MeV)
+ 22.4
(10) D + 6Li → 2  4He
MeV
(1.7 (2.3
(11) p + 6Li →   4He +   3He
MeV) MeV)
+ 16.9
(12) 3He + 6Li → 2  4He   +  p
MeV
+ 8.7
(13) p + 11B → 3  4He
MeV

p (proton), D (deuterium), and T (tritium) are shorthand notation for the first three
isotopes of hydrogen. For reactions with two products, the energy is divided between
them in inverse proportion to their masses, as shown. In most reactions with three
products, the distribution of energy varies. For reactions that can result in more than
one set of products, the branching ratios are given.
Some reaction candidates can be eliminated at once.
[1] (http://theses.mit.edu/Dienst/UI/2.0/Page/0018.mit.theses/1995-130/30?
npages=306) The D-6Li reaction has no advantage compared to p-11B because it is
roughly as difficult to burn but produces substantially more neutrons. There is also a
p-7Li reaction, but the cross section is far too low except possible for Ti > 1 MeV, but
at such high temperatures an endothermic, direct neutron-producing reaction also
becomes very significant. Finally there is also a p-9Be reaction, which is not only
difficult to burn, but 9Be can be easily induced to split into two alphas and a neutron.
In addition to the fusion reactions, the following reactions with neutrons are important
in order to "breed" tritium in "dry" fusion bombs and some proposed fusion reactors:
n + 6Li → T + 4He
n + 7Li → T + 4He + n
To evaluate the usefulness of these reactions, in addition to the reactants, the products,
and the energy released, one needs to know something about the cross section. Any
given fusion device will have a maximum plasma pressure that it can sustain, and an
economical device will always operate near this maximum. Given this pressure, the
largest fusion output is obtained when the temperature is chosen so that <σv>/T² is a
maximum. This is also the temperature at which the value of the triple product nTτ
required for ignition is a minimum. This optimum temperature and the value of
<σv>/T² at that temperature is given for a few of these reactions in the following
table.

fuel T [keV] <σv>/T² [m³/s/keV²]

D-T 13.6 1.24×10-24

D-D 15 1.28×10-26

D-3He 58 2.24×10-26

p-6Li 66 1.46×10-27

p-11B 123 3.01×10-27

Note that many of the reactions form chains. For instance, a reactor fueled with T and
3He will create some D, which is then possible to use in the D + 3He reaction if the
energies are "right". An elegant idea is to combine the reactions (11) and (12). The
3He from reaction (11) can react with 6Li in reaction (12) before completely
thermalizing. This produces an energetic proton which in turn undergoes reaction (11)
before thermalizing. A detailed analysis shows that this idea will not really work well,
but it is a good example of a case where the usual assumption of a Maxwellian plasma
in not appropriate.

You might also like