Sie sind auf Seite 1von 13

How a Centrifuge Works

Centrifuge Basics
by Ivan Oelrich and Ivanka Barzashka
Uranium powers both nuclear reactors and nuclear bombs. However, uranium cannot be used in its natural form-- it must be processed to increase the
concentration of the active isotope of uranium, U-235. Most chemical elements have different isotopes, which are atoms with slightly different weights
resulting from the different number of neutrons in the nucleus. While the chemical properties of two isotopes are almost identical, the difference in the
number of neutrons in the nucleus can result in nuclear properties that are dramatically different. Such is the case with uranium, which is made up of two
isotopes, the predominant uranium-238 (or U-238) and U-235, which is only 0.7% of natural uranium. It is the U-235 that can be directly split, or
fissioned, to produce power in most of the worlds current commercial nuclear reactors. (Thus, materials like U-235 and plutonium are called
fissionable".) The U-238 is not directly fissionable. To be useful, the concentration of U-235 must be increased to 3-5 percent for a typical commercial
reactor and to 80-95 percent for a nuclear weapon.
Because the two isotopes of uranium are chemically indistinguishable, physical methods that exploit the small difference in weight must be used to enrich
uranium in the 235 isotope. Historically, several techniques have been used but the method that is overwhelming preferred today is a gas centrifuge.
Surprisingly, for such a heavy metal, uranium can form a compound that is a gas at moderate temperatures. Uranium combines with six atoms of fluorine
to form uranium hexafluoride (UF6). At room temperature and pressure, UF
6
is a white solid but it turns to a vapor or gas at moderate temperature (133 F
or 56 C at normal atmospheric pressure).
Coincidentally, fluorine has only one stable isotope, F-19. This is important because, if the fluorine atoms had different weights, there would be no way to
distinguish whether the difference in weight of the UF
6
molecule was due to the uranium or the fluorine atoms. Since fluorine atoms are identical, any change in mass has to be due to the different
uranium isotopes.
How a CentrifugeWorks
Sections

Centrifuge Basics
General description of what a centrifuge is
and the principles that govern its work,
explanation of large-scale enrichment by
cascades and nuclear weapons proliferation
concerns
Engineering Considerations
Technical discussion of centrifuge
components and the challenges in building
and running a gas centrifuge
Separation Theory
An overview of the theory of isotope
separation in the generalized case of a
separating unit.

The heart of a gas centrifuge is a tube, called the rotor, that spins at high speed around its long axis. The performance of the centrifuge depends critically on the speed of
the rotor. Technical advances in materials, high-speed bearings, and precision machining are what have made increased rotor speeds possible and made centrifuges
practical. Today, rotors may spin in excess of 60,000 rpm with the outside surface of the rotor moving well in excess of the speed of sound.
The spinning rotor creates powerful centrifugal forces that mimic a miniature gravitational field except, of course, up is toward the axis of the rotor and down is
toward the outer rim. The Earths atmosphere is densest at the surface and, as altitude increases, each layer of air has less air above it pressing down on it and compressing
it. Hence, the pressure and resulting density decrease as we go up through the atmosphere. A similar effect occurs in a centrifuge but, because the forces created by the fast
spinning centrifuge might be a million times stronger than gravity, everything happens on a much smaller scale.
In the spinning rotor, the slightly heavier UF
6
that contains U-238 will be slightly more compressed along the rim relative to the lighter UF
6
that contains U-235, which
would have a relatively greater concentration near the axis. The separation between the two isotopes created by the centrifugal forces is quite small. However, there is a
simple way to substantially increase the degree of separation by creating a circulation flow along the length of the centrifuge. If one end of the centrifuge is heated, the
warmer gas will rise at that end and flow toward the opposite end along the axis, while cooler gas will flow in along the wall to replace it. Alternately, a scoop used for
removing gas at one end will slow the flow of gas, reducing the centrifugal force and allowing the gas to rise. Now the centrifuge has a flow of gas along the center toward
one end and a flow along the wall toward the other end.
As in any gas, the molecules of UF
6
are not static, rather individual molecules have thermal energy and are moving around within the volume of gas. To say that the
concentration of U-238 is higher along the wall and U-235 higher along the center, simply means that, statistically, each molecule containing 238 will spend more time on
average along the wall than a molecule containing U-235, which will, statistically, spend a relatively longer time near the center. Being more often near the wall, the U-
238 will get pushed along with the wall flow in the direction of the hot or scoop end of the rotor more often than it will get pushed along the center flow, thereby
increasing its concentration at that end. The U-235 will spend, on average, more time being pushed along with the flow along the center and will concentrate at the opposite end. Thus, the U-235 and
U-238 do not separate relative to the wall and the center but along the length of the centrifuge. This is one reason that the longer the rotor, the greater the separation of the centrifuge. Tubes running
along the center of the rotor extract the UF6 gas. From one end the gas will be slightly enriched in U-235 compared to the feed gas and at the opposite end the gas will be slightly depleted in U-235
compared to the feed gas.
One of the key determinants of a centrifuges performance is the speed of rotation and this depends on the material the rotor is made of. Because only a small amount of UF
6
is in the centrifuge at any
given moment, most of the force on the spinning centrifuge is due to the centrifuge material itself. The important characteristic of the centrifuge material is, therefore, not just its strength but the ratio
of its strength and density. For example, two common materials used for centrifuge rotors are aluminum and steel. Aluminum is only about a quarter as strong as steel but that lesser strength is
compensated by a density that is a third lower, so the much weaker, but lighter, aluminum is able to spin almost as fast as the stronger, but heavier, steel.
High spin rates require very precise balance of the rotor so another critical technology is high precision machining of the rotor components. High speed, low wear and low friction bearing are also
essential. Finally, UF
6
is highly corrosive so all parts that are exposed to the gas must be made of resistant materials.

EG-1802 Soviet
Centrifuge

No single centrifuge can achieve an adequate degree of enrichment and the amount of material that a single centrifuge can handle is small. To process sufficient
material, centrifuges, sometimes hundreds, are operated in parallel. To get the necessary degree of enrichment, the enriched output of one set of centrifuges will be
fed in as input to another set of centrifuges for further enrichment. Each set of centrifuges enriches the uranium a bit more than the previous until the desired enrichment is achieved. Such a collection
of centrifuges is called a cascade and any given set of centrifuges operating in parallel is called a stage. The other part of the output of the first stage will, of course, be slightly depleted in U-235. It
will, nonetheless, still contain significant quantities of the valuable isotope, so it is not simply thrown away. It is fed into centrifuges that enrich the depleted material back up to the original
concentration and then fed back into the enriching cascade. The stages above the initial feed point that enrich the uranium above its original concentration are the enriching stages and those that
operate below the original concentration are called stripping stages.
A cascade to supply a large nuclear reactor may require several thousand centrifuges of moderate performance. Depending on the enrichment of each stage and the desired degree of enrichment, a
cascade may have anywhere from a few stages to more than twenty.
Centrifuges raise serious nuclear weapons proliferation concerns because exactly the same machines that are used to enrich uranium for a nuclear reactor can enrich uranium for a nuclear bomb. In
general, a nuclear reactor needs a small degree of enrichment of a large amount of material and a bomb needs a large degree of enrichment of a small amount of material. Exactly the same centrifuges
can do either job; the only change required is how they are piped together into a cascade. Moreover, a single typical large commercial nuclear power plant may have ten times more separative work
than is needed to produce one uranium bomb per year, so even a modest commercial enrichment facility has a significant nuclear weapons production capability.

Engineering Considerations for Gas Centrifuges
by Ivanka Barzashka and Ivan Oelrich
Establishing a domestic centrifuge program starts with developing technical expertise and acquiring or creating centrifuge engineering designs. Once this has been
achieved, the construction phase includes obtaining the materials, manufacturing the centrifuge components, assembling and balancing the machines. The machines
are then tested and accelerated up to their maximum operational speed. Once a centrifuge is running, the controllable variables of the machine (such as throughput
and cut) have to be optimized to maximize its separative capacity. Ultimately, individual machines are piped together in cascades, which are designed for specific
enrichment and throughput.
This article describes how centrifuges are built and how they are run. It discusses the engineering aspects of the main centrifuge components and stresses the
challenges of building and operating the machines. First, the choice of material for the manufacture of centrifuge components is discussed. Then, we look at the
major rotating components, the feed and extraction system, the casing and the power supply.
Materials for Building a Gas Centrifuge
The first challenge in developing a gas centrifuge program is obtaining the materials to manufacture the required parts. High strength, corrosion resistant material
for the rotors, baffles, and bellows, as well as homogeneous magnets for top bearings may not be possible to manufacture domestically. However, proliferation-
sensitive materials are included in a dual-use items list prepared by the International Atomic Energy Agency (IAEA) and their export to non-nuclear-weapon states is monitored [6].
Tensile strength and density are the most significant characteristic of the rotating centrifuge components and especially the rotor. The ultimate tensile strength characterizes the maximum stress that
a material can endure before it ruptures. The ratio of the tensile strength to the density of the material determines the maximum peripheral speed v of the rotor.
Urenco Centrifuge Cascade

Note that the peripheral speed is not the rotational speed, but is the rotational frequency multiplied by the circumference of the rotor. A large radius and a low rotational speed can have the same
peripheral speed as a small radius and a high rotational speed.
The elastic modulus, or the tendency to deform elastically under pressure, is a measurement of the materials stiffness and closely correlates with the tensile strength. It is important in considering
rotor material because it determines the flexural resonances of the rotating cylinder (discussed in more detail in the section below), which, in turn, determine the critical speeds of the rotor.
A centrifuge is required to spin continuously for long periods of time. Material properties such as creep and fatigue are considered to predict component and machine wear caused by periodic stress
that is below the threshold to cause immediate failure.
The mass and stiffness of the material directly affect the centrifuges maximum speed. In turn, the overall separation performance is sensitive to maximum speed. Theoretically, the maximum degree
of separation of the centrifuge is proportional to the peripheral speed raised to the fourth power. Thus, a 10 percent increase in the peripheral velocity would result in a 46 percent increase in
maximum separation. In practice, engineering tradeoffs and inevitable inefficiencies in the machine make the increase substantially smaller, with some authors suggesting that separation proportional
to the speed squared is a more reasonable approximation [8].
Early centrifuges used rotors made from high-strength aluminum alloys. More advanced designs use higher tensile strength materials with such as maraging steel,
titanium, or glass or carbon fiber, resulting in significant increases in speed. Values range from 400 m/s for aluminum alloys, 440 m/s for titanium, 500 m/s for
maraging steel, 522 m/s for glass fiber, and 720 m/s for carbon fiber. [3]

Since uranium hexafluoride (UF
6
) is a corrosive gas, the interior surfaces of the centrifuge need to be resistant to attack [2] and free of organic material, such as
grease, hydrocarbon oil from bearings or even human fingerprint oil. This includes not only rotating components such as the rotor, but stationary ones such as feed
and extraction piping.
Some materials are essentially inert to UF
6
[15] because a thin film of metal fluoride forms, which protects the bulk of the material from further
corrosion [16],[12]. According to the IAEA trigger list, stainless steel, aluminum, nickel and alloys containing more than 60 percent nickel are resistant to corrosion and are monitored. Copper and
copper alloys such as bronze are also UF6 resistant [15], but are not the best choice for a centrifuge rotor due to relatively higher density.
Metals are sometimes anodidized, that is pre-oxidized, to improve corrosion resistance. This is usually done to maraging steel rotors using controlled atmosphere furnaces using a mixture of air and
steam to control oxidation [14].
Constructing Centrifuge Components
Rotor
Constructing a centrifuge rotor that is perfectly balanced, that is, a rotor in which the geometric and mass axes coincide, is practically impossible. The difference between the two axes creates a
rotating radial load a net force that pulls the rotor laterally. Mass imbalance in a single plane can be detected when the rotor is at rest. Mass can also be out of balance at different points along the
length of the rotor, creating forces that cause rocking of the rotor axis. This is called a dynamic unbalance because it cannot be detected without spinning the rotor. At high speeds, the unbalanced
force can rock the rotor within its bearings, creating vibrations.
Unbalance forces can create synchronous whirls, which are rotations about the axis formed by the bearings occurring at the same frequency as the rotor rotation. The
cylinder tries to rotate about its center of mass but is being pushed toward the axis defined by the bearings and their geometric center. There are two types of whirling
in centrifuge rotors cylindrical (or translatory [9]) and conical. The former is a horizontal displacement perpendicular to the axis of rotation and the latter is a tilt
about the axis.
A rigid rotor has an effective mass and is supported by bearings, which provide a restoring force and act like springs. Any system with an inertial mass and restoring
force or stiffness will have natural frequencies. When an external driving force appears, such as the unbalance force, and it has the same frequency as one of the
natural frequencies, resonance is created. In centrifuges, these resonances are called critical speeds.
If we ignore the flexibility of the rotor and just consider the rotor mass and the stiffness of the bearings, these critical frequencies are called the rigid-body critical speeds. There are two types of
rigid-body critical speeds due to the two types of whirls. The cylindrical and conical critical speeds are given respectively by the equations below, where S is the stiffness of the bearings, M is the
mass of the rotor, l is the half bearing separation, I is the transverse inertia and P is the polar inertia [17].


Rotors can be meticulously balanced, which can greatly decrease the unbalance force but cannot eliminate it completely. Vibration can be mitigated by design of a flexible shaft-bearing system to
allow the rotor to locate and freely spin around its center of gravity. Lowering the bearing stiffness also lowers the natural frequency making the resonance occur at a lower speed and reducing the
total load because load is the displacement times the speed [17].
The ratio of the transverse to polar inertia increases with length and the conical critical speed is reduced, thus, for a longer rotor, traversing the rigid-body criticals becomes trivial [17]. However, a
longer rotor encounters flexural critical speeds, which occur at higher frequencies. Whereas the rigid-body criticals depend on the vibrations within the bearings, the natural frequencies of flexure are
due to longitudinal vibrations of the rotor itself (because we consider the bearings to be rigid), which depend on the modulus of elasticity E and the density of its material . At certain speeds, the
frequency of rotation equals the natural frequencies and resonance occurs. During resonance, the rotor will bend in the shape of the harmonics: for the first critical speed it will take the shape of an
arc, for the second like an S and so forth. The critical flexural frequency is given by the equation below, where L is the length of the rotor, r is its radius, E is the modulus of elasticity and is the
density, is an eigenvalue and i is the mode number[3]:


There are two types of centrifuges: subcritical and supercritical. A rotor is subcritical when it operates at a rotational frequency less than the lowest flexural resonance frequency and supercritical
when operates above it. The flexural critical conditions can be expressed in terms of the ratio between the rotors length and its radius. Therefore, knowing the dimensions of the rotor and its material
properties, we can determine whether it is subcritical or supercritical. A rotor has many resonant frequencies and passing through each one creates a mechanical risk to the centrifuge. The goal is to
maximize the separative capacity before hitting the next higher critical frequency, so centrifuges are usually designed to operate just below a critical frequency.
Traversing the flexural critical speeds is the major obstacle in getting supercritical centrifuges to rotate at their optimum speeds. There are three ways to do this: accelerate quickly through the critical
speed, optimize the damping on the bearings, or insert bellows.
By quickly accelerating the rotor through the critical speeds, there will be no time for the amplitudes to reach destructive levels. This method was used with the Manhattan supercritical centrifuge at
the University of Virginia in the 1940s. A powerful motor is required to drive the centrifuge rotor for this option. For industrial purposes, a more energy-efficient approach is to optimize the
dampers, which is not a simple task. The problem of vibration damping was one of the reasons gas centrifuge were abandoned by the U.S. during the World War II Manhattan Project [17].
Bellows
Another way to help the supercritical centrifuge reach its maximum speed is to introduce bellows (or sylphons) along the length of the rotor. These are flexible joints that break up the rotor into
several parts. Bellows have two benefits: they allow the rotor to bend and take up the shapes of the harmonics without breaking and they reduce the resonance frequencies, thereby decreasing the total
force when the rotor passes through the critical speeds. Careful design and placement of the bellows allows management of the resonant frequencies. Bellows were first used in Soviet Union
supercritical machines during the late 1940s and early 1950s [13].
Bearings
The purpose of bearings is to constrain a rotating object to spin about a specific position. Bearings behave like springs, since they apply a restoring force to set the rotating
object back to its original position. The spring force constant or, in the case of bearings, the stiffness, is therefore an important parameter to consider. Two types of bearings
were used for gas centrifuges. Hydrodynamic journal bearings were used in the Groth and Beams machines. This type of bearings is characterized by high stiffness and high
power consumption. Zippe machines used a combination of a pivot and a magnetic bearing [3]. The magnetic bearing consisted of annular-shaped permanent magnets and was
placed on the top of the centrifuge. It was used to relive the stress on the bottom bearing. Magnetic bearings have very little power loss and are suitable to work in a vacuum.
This greatly reduces wear and energy consumption but also greatly lowers the stiffness of the bearing. Lower stiffness reduces the natural frequency, which diminishes the
radial load caused by unbalancing. However, this also makes the centrifuge more delicate and susceptible to perturbations [17].
Extraction System: Scoops and Baffles
Material is extracted from inside the rotor by small stationary tubes called scoops. The scoops are located at the top and bottom of the rotor and fixed to the central gas extraction system. They
extend out to the periphery of the rotor, since this is where most of the material is located due to the centrifugal forces created by the rapid spin. The scoop, which ends in a hook-shaped Pitot tube,
faces the circulating gas and pressure is created, which pushes up the material. No extra power is required to move the gas to the next machine in a cascade because the scoops use the momentum of
the rotating gas and the high pressure at the periphery. In a machine with thermally-driven circulation, the warm end scoop extracts the heavier material more abundant in U-238 and the cool-end
scoop the lighter material richer in U-235. Sometimes scoops are shielded by baffles, so that no disturbance in the countercurrent is created. The baffles are rotating perforated discs, which separate
the part of the rotor where separation takes place from the part of the rotor where the material is extracted. Unbaffled scoops can contribute to the countercurrent flow inside the machine, being the
equivalent of heating. Aerodynamic properties of the scoops have an effect on the separation factor of the centrifuge and should be optimized for best performance. Scoops are made from corrosion
resistant materials, but do not need to have the tensile strength of the rotor because they are stationary.
Vacuum Casing
The outside surface of a modern centrifuge rotor is traveling well above the speed of sound so centrifuges are enclosed in a vacuum casing to minimize drag. The casing is also
designed to contain an exploding rotor [2]. Operating the rotor in a vacuum also makes rotor temperature control easier, eliminating any unwanted convection in the gas in the rotor[16]. In addition,
the casing isolates the machine from outside vibration, particularly from the other machines in the cascade.
Although the rotor shafts are sealed, it is possible for small amounts of UF
6
to escape. A molecular pump is used to maintain the vacuum inside the centrifuge casing. Molecular pumps used in
centrifuges are a type of drag pump. They consist of a smooth cylinder made of UF
6
resistant material with spiral grooves, which is attached to the inside of the casing close to the rotor [6]. The
movement of the rotor imparts a momentum to molecules that hit it and move them in the direction of rotation, or "drags" them along. The spiral grooves channel the molecules away from the low
pressure end to the high pressure end. The most advanced type of drag pump is the Holweck pumps, which was what was used in the Zippe centrifuge [17].
Power Supply
The motor drive is operated in low voltage in high vacuum. The motor stators need to be especially designed for synchronous operation in vacuum. The frequency
changers control the rotational speed of the rotors by supplying the rotor stators with power at the correct frequency. They need to have high stability, low harmonic
distortion and high efficiency because they need to be able to successfully accelerate the rotors above critical speeds and avoid overheating in vacuum. Frequency changers
convert the alternating current from the power grid to a higher frequency ranging from 600 Hz to 2000 Hz [6].
Ancillary Equipment
Ancillary, test, and production equipment such as such variable frequency power supplies for the drive motors, vacuum pumps, precision mass spectrometers or vibrational
test systems can be obtained off-the-shelf. However, proliferation-sensitive materials, parts, and technology are included in IAEA's trigger list and their international trade
is monitored.
Designing and building gas centrifuges pose significant engineering challenges. Some technological difficulties can be overcome in the laboratory but the solutions may
prove to be uneconomic for large-scale production. For example, the United States developed prototype supercritical centrifuges but chose gaseous diffusion as the industrial method for uranium
enrichment for its World War II Manhattan Project. Historically, the least time required to develop indigenous centrifuge technology has been 8 years, in the case of the Soviet Union. Even using
acquired Urenco designs, it took Pakistan 6 years for the development of a centrifuge program, designed specifically for the production of bomb-grade uranium [18].
We would like to thank David Chichka, assistant professor of Mechanical and Aerospace Engineering at George Washington University, for his contribution and insights.
References
[1] Adams, M. L. (2000). Rotating Machinery Vibration: From Analysis to Troubleshooting . CRC.
[2] Avery, D. G., & Davies, E. (1973). Uranium enrichment by gas centrifuge. London: Mills and Boon.
[3] Benedict, M., Pigford, T. H., & Levi, H. W. (1981). Nuclear chemical engineering (2nd ed.). New York: McGraw-Hill.
[4] Callister, W. D. (2003). Materials Science and Engineering: An Introduction (6th ed.). New York: John Wiley & Sons, Inc.
[5] Den Hartog, J. P. (1985). Mechanical Vibrations (4th ed.). New York: Dover Publications.
[6] International Atomic Energy Agency. (2006, March 20). INFCIRC/254/Rev.8/Part 1a. Retrieved April 2, 2009,
from http://www.iaea.org/Publications/Documents/Infcircs/2006/infcirc254r8p1.pdf
[7] Jousten, K. (2008). Handbook of Vacuum Technology (B. Nakhosteen, Trans.). Wiley-VCH.
[8] Kemp, R. S. (2009). Gas Centrifuge Theory and Development: A Review of U.S. Programs. Science and Global Security, 17(1).
[9] Rao, J. S. (1996). Rotor Dynamics (3rd ed.). New Age International.
[10] Redhead, P. A. (1997). Vacuum Science and Technology Pioneers of the 20th Century. New York: American Institute of Physics.
[11] Thomson, W. T., & Dahlen, M. D. (1998). Theory of Vibrations with Applications (5th ed.). Upper Saddle River, NJ: Prentice Hall.
[12] United States Enrichment Corporation. (1995, January). Uranium Hexafluoride: A Manual of Good Handling Practices (USEC-651, Revision 7). Retrieved April 2, 2009,
fromhttp://web.ead.anl.gov/uranium/guide/ucompound/propertiesu/uranium.cfm
[13] U.S., Central Intelligence Agency. (1957, October 8). The Problem of Uranium Isotope Separation by Means of Ultracentrifuge in the USSR (EG 1802). Retrieved April 2, 2009,
fromhttp://www.fas.org/irp/cia/product/zippe.pdf
[14] U.S. Department of Energy, Nuclear Transfer and Supplier Policy Division. (1996). Handbook for Notification of Exports to Iraq: Annex 3 (Vol. SARC - 001/98). Retrieved April 2, 2009,
fromhttp://www.iraqwatch.org/government/US/DOE/DOE-Annex3.htm
[15] U.S. Department of Energy. (1999, April). Final Programmatic Environmental Impact Statement for Alternative Strategies for the Long-Term Management and Use of Depleted Uranium
Hexafluoride (DOE/EIS-0269), Appendix A: Chemical Forms and Properties of Uranium. Retrieved April 2, 2009, from http://web.ead.anl.gov/uranium/pdf/app_a.pdf
[16] Villani, S. (1976). Isotope Separation. American Nuclear Society.
[17] Whitley, S. (1984). Review of the gas centrifuge until 1962. Part II: Principles of high-speed rotation. Reviews of Modern Physics, 56(1), 67-97.
[18] Zentner, M. D., Coles, G. L., & Talbert, R. J. (2005). Nuclear Proliferation Technology Trends Analysis (Tech. No. PNNL-14480). Pacific Northwest National Laboratory.

Separation Theory
by Ivanka Barzashka and Ivan Oelrich
The following discussion ignores the internal mechanism of how a centrifuge works and treats the machine as a black box. To understand the separative
properties of a centrifuge, only material flows in and out of the machine will be considered. Assuming that no material is created or lost, we can track the flow of mass and describe the degree to
which the concentration of the desired isotope is increased.
For any separating element, a stream of material of certain concentration enters and two streams of material emerge one with higher and the other with lower concentration of the desired isotope
than the input. The input material is called the feed. The output stream with high concentration is called the product (or heads) and the stream with low concentration the waste (or tails). The
general principles that describe the isotope separation process are the same for any single separating element or multiple elements linked together to function as a single unit, called a cascade. A
detailed discussion of cascade theory is available in the section on Cascade Theory.
Input Material for a Gas Centrifuge
The material that enters and leaves a gas centrifuge is uranium hexafluoride (commonly referred to as hex or UF
6
), a gaseous compound of uranium. The uranium in UF
6
is a mixture of primarily two
isotopes, U-238 and U-235. Fortuitously, the fluorine has only one isotope, so any mass difference between molecules of UF
6
is due to the uranium. For uranium enrichment, the isotopic composition
of interest is the amount of U-235 present with respect to the total mass of the substance, mostly U-238. Natural uranium contains a third isotope, U-234, but typically in concentrations of 50 to 60
parts per million. U-234 is typically ignored in mathematical discussions of enrichment and uranium is considered simply as a two component system. (U-234 concentrations in natural uranium vary
slightly from mine to mine and these small differences in U-234 concentration can help identify the source of illicit nuclear material. Such tracking is one example of nuclear forensics.)
In a gas centrifuge used for uranium enrichment, the feed is typically natural uranium with a composition of 0.711 percent U-235. In a batch recycling process, it is possible to use low-enriched
uranium (LEU), containing 3.0 to 5.0 percent U-235 as feed to produce bomb-grade uranium.
When flow rates or total amounts are reported, care should be taken to note whether the quantities are measured in terms of uranium mass or UF
6
mass. The mass of uranium hexafluoride (UF
6
) is
equal to 1 uranium atom and 6 fluoride atoms. The molecular weight is therefore given by:

This makes the mass of uranium approximately 67.6 percent of that of uranium hexafluoride. IAEA reports usually describe material feed rates in terms of UF
6
. However, flow rates should be
converted to uranium when calculating separative properties such as the separative power.
Defining Physical Concepts
Before describing how to quantify the separation capability of a centrifuge, several physical terms must be defined.
The relative concentration of the isotope that is being enriched is called the isotopic composition and measured as a fraction or percent of the total mass. The isotopic composition of the feed is x
f
.
Neglecting the U-234, the concentration of the U-238 in the feed will be 1-x
f
.
The feed goes into the separating unit at a flow rate or throughput of F. For an individual machine, throughput is usually measured in grams or kilograms per hour and for a cascade in tons per year.
The product stream has concentration x
p
and flow rate P. The waste has a concentration and flow rate of x
w
and W.
The cut of the separating element is defined by the ratio between the product and feed flow:

Since no material is lost in the separation process, the amount of U-235 coming into the separating unit as feed should equal the amount of U-235 coming out as product and waste. The material
balanceequation relates the compositions and flow rates of the feed, product and waste.

In terms of the cut, the material balance equation yields:

The relative isotopic abundance R is another parameter that describes the composition of the material. It relates the quantity of the two isotopes in the mixture by taking the ratio of the desired
isotope (U-235) to the undesired one (U-238). For example, the relative isotopic abundance of the waste is given by:

Separation Factor
The separation factor (also called single stage separation factor or simple process factor) measures the degree of separation achieved by a separation element. It is defined as the ratio between the
relative isotopic abundance of the product and that of the waste.

Because the separation factor is often close to unity, the separation gain (also called the simple process difference) is sometimes useful.

For separation through gaseous diffusion, each discrete separative element enriches the material only slightly and the separation factor is close to one. When single units are interconnected in a
cascade, the overall separation factor of the series of machines is the product of all the individual units separation factors multiplied together.
In general, there is a tradeoff between separation factor and throughput: during a fixed time, a machine can enrich a large amount of material to a small degree or a small amount of material to a large
degree. The greatest separation occurs at total reflux i.e. letting the centrifuge run with a fixed charge of UF
6
but without adding or removing any material. Clearly, this does not produce any
separated product. At the opposite extreme, quickly running large quantities of UF
6
through the machine produces a lot of product, but the centrifuge does not have time to work on the mixture to
achieve separation. The objective in operating a separative device is to balance the separation and throughput to maximize the overall performance.
Separative Work and Separative Power
We need a way to define and measure the overall effect and capacity of a centrifuge or other enriching element. For that purpose, a separative work and a separative power have been defined.
Theseparative work refers to a certain enhancement of enrichment on a certain amount of material. The separative power describes the rate at which the separative work is done on that material.
The degree of enrichment produced on a mass of material by any particular machine depends on the composition of the feed. No enrichment takes place, obviously, if a centrifuge is working on an
isotopically-pure gas, even though the machine is operating exactly in the same way as it would if it were filled with a gas made of two isotopes. A measure of effectiveness of the machine should,
therefore, be independent of the composition of the feed material and is more complex than simply the degree of enrichment produced by a machine with a particular feed material.
Paul Dirac first treated this problem in 1941 and developed the notion of the value of an enriched gas, which closely correlates to entropy. The performance of the enriching device is then defined in
terms of net value added to a given amount of the material that passed through the machine for enrichment. The value of a gas is simply its mass times a value function V(x). The separative
work U of a centrifuge is the net gain in value, V, which is the value of the output (the value of the product plus the value of the waste) minus the value of the input (feed material).

The value function can be derived either from the change in entropy for the gas [3] or from the general equation for separative work (above).The latter mathematically defines the value function so
that the separative work of an enrichment device is independent of concentration. The value function can be graphed on a simple plot or can easily be calculated for a given concentration of material.
The equation for the value function is:

where x is the concentration of the material, which can be that of either the feed, product or waste. (Actually, in the derivation, the concentrations of the feed, product, and waste come into the
equation, but under the close-approximation condition, it can be assumed they are all close enough to be considered equal.)
Note that the value is a dimensionless number and a thermodynamic intensive quantity. Increasing the value by 1 is called a separative work unit, or SWU. Separative work is measured in units of
kgSWU, which is the increase of value by 1 of a kilogram of material. Since almost all references to separative work units are kgSWUs, the kg is sometimes dropped and the unit is called, simply, a
SWU. Very early texts from Britain and the United States will sometimes refer to poundSWUs and the separative work of an entire enrichment plant will be described in units of ton-SWUs, so it is
best to always include the mass unit when using SWUs and explicitly refer to kgSWUs. Note that the numeric result for the value of any given material is not really important, what measures the
effectiveness of the centrifuge is the net change in value. It turns out that using the value function equation above, the value of an equal mixture of isotopes is zero.
Following the analogy between work and power with power being work per unit time, the separative power is simply the separative work done in a fixed time. Separative power is usually expressed
in units of kgSWU/hr for individual machines and tonSWU/year for entire enrichment factories.
The value as a function of concentration is shown in the figure adjacent. Note that the separative work required for any fixed, or even relative, increase in
concentration depends on the concentration, as was hypothesized above. So a 10 percent increase from 0.10 to 0.11 U-235 concentration takes more separative work than the 10 percent increase
going from 0.50 to 0.55. (Beyond a concentration of 0.50, concentration of U-235 becomes increasingly difficult because U-235 becomes the majority isotope and U-235 and U-238 simply trade
places mathematically.)
Knowing the value added by a given machine, one can quickly see how much enrichment occurs at any given feed concentration. Conversely, one can calculate the amount of value that must be
added to achieve a certain enrichment of a given amount of material and then calculate how much centrifuge work would be required to produce that increased value. Enrichment value allows a direct
comparison of the amount of separative work required to produce, for example, a ton of 5 percent U-235 for a nuclear reactor and 50 kg of 95 percent U-235 for a nuclear bomb.
Examples
We can calculate the amount of separative work needed to produce enough LEU to power a pressurized water reactor (PWR) with a 1000 MWe (megawatt electric) gross capacity, which is the
electric power that the reactor produces. Factoring in the inefficiency of the PWR, gives a thermal capacity (or the energy that needs to be produced in fission) of 3250 MWt (megawatt thermal). An
average reactor has a capacity factor of 0.8, which means that it produced about 80 percent of the maximum net electricity that it could have generated at constant full-capacity operation. Current U.S.
reactors have an average capacity of about 90 percent.
Although most of the energy in the reactor is produced from fission of U-235, there is some energy produced due to internal breeding. An accurate calculation of
the amount of fuel needed to power a reactor requires the use of a computer program. According to one source, a reactor with the above characteristics with a burnup of 33,000 MWd/MT (megawatt
days per ton of fuel) and three zone fueling requires about 27,300 kg uranium containing 3.3 percent U-235 [2]. Not all of the U-235 is burned in the reactor. There is less than 1 percent present when
the fuel rods are removed from the reactor core.
If fuel is not recycled, factoring in a 1 percent loss during conversion and fuel fabrication, a capacity of 109 t SWU is needed to produce 27,500 kg LEU. This is valid assuming we are enriching
natural uranium starting at a concentration of 0.711 percent and enriching to 3.3 percent LEU with tails of 0.3 percent.
To enrich an IAEA significant quantity of HEU (equivalent to 25 kg U-235 or 27.8 kg HEU at 90 percent concentration) from natural uranium with a tails of 0.3 percent requires 5360 kg SWU.
This means that if we use the separative capacity of a plant sized to produce enough LEU for a reactor, we could produce 565 kg of HEU or enough for 20 nuclear bombs. In this scenario, we would
be throwing away 42 percent of the U-235 present in the feed.
If fuel grade uranium (LEU at 3.3 percent) is further enriched to get HEU and the tails are set at 0.3 percent, the separative capacity of the enrichment plant would allow for the production of enough
material for 52 bombs in one year. About 7.5 percent of the U-235 present in the feed (in this case LEU) will be discarded in the process. If the same scenario is used, but with a tails of 1.2 percent, 30
percent of the U-235 present in the feed will be discarded, but the HEU produced in a year will be enough for 85 bombs. Note that in these scenarios, LEU is not the limiting quantity.
References
[1] Olander, D. R. (1972). Technical Basis of the Gas Centrifuge. Advances in Nuclear Science and Technology, 6, 105-165.
[2] Pigford, T. H., & Levi, H. W. (1981). Fuel Cycles for Nuclear Reactors. In M. Benedict (Author), Nuclear chemical engineering (2nd ed., pp. 84-156). New York: McGraw-Hill.
[3] Whitley, S. (1984). Review of the gas centrifuge until 1962. Part I: Principles of separation physics. Reviews of Modern Physics, 56(1), 41-66.

Das könnte Ihnen auch gefallen