Download as pdf or txt
Download as pdf or txt
You are on page 1of 21

Chapter 10

Cryogenic Systems

Thomas J. Dolan

Objectives

After reading this chapter one should understand


Materials properties at low temperatures
Cryogenic refrigeration
Insulation and cryostats
ITER cryogenic systems.

10.1 Introduction

The term ‘‘cryogen’’ was originated by H. K. Onnes from the Greek stems ‘‘cryo’’,
which means ‘‘cold,’’ and ‘‘genes,’’ meaning ‘‘that which generates.’’ Today, the
term cryogenics refers to the physical, chemical, engineering, and industrial
applications of phenomena at low temperatures, below about 123 K (-150 C).
The four common temperature scales are related by the equations

TðFÞ ¼ 1:8 Tð CÞ þ 32


TðRÞ ¼ TðFÞ þ 459:67 ð10:lÞ

TðKÞ ¼ Tð CÞ þ 273:15
Study of cryogenic phenomena is roughly one hundred years old. Early
developments in production of liquid oxygen and nitrogen occurred in Europe in
1877–1890. James Dewar invented vacuum insulated flasks in 1892, and he liq-
uefied hydrogen in 1898. Carl Von Linde and Georges Claude developed practical
systems for the liquefaction of air in 1895 and 1902. H. K. Onnes liquefied helium

T. J. Dolan (&)
NPRE Department, University of Illinois, Urbana, IL 61801, USA
e-mail: dolantj@illinois.edu

T. J. Dolan (ed.), Magnetic Fusion Technology, Lecture Notes in Energy 19, 491
DOI: 10.1007/978-1-4471-5556-0_10,  Springer-Verlag London 2013
492 T. J. Dolan

(the last gas to be liquefied) in 1908 and discovered superconductivity in 1911.


After the end of the First World War, production of industrial gases (such as
oxygen and nitrogen) by cryogenic processes grew rapidly.
The first expansion engine for liquefying helium was built by P. Kapitza in
1934. A comparatively economical process for producing liquid helium was
developed by S. C. Collins in 1947, which facilitated expansion of research on
low-temperature phenomena.
Large V-2 rockets fueled by liquid oxygen and alcohol were developed during the
Second World War. Development of rocket technology continued in the USSR and in
the USA, leading to the launching of satellites and space flights. Cryogenics has now
become an important industrial field with many practical applications, Table 10.1.
Cryogenic systems are essential to the success of neutral beam injection sys-
tems and superconducting magnets. Cryopumps, having the highest attainable
pumping speed, prevent excessive influx of cold neutral gas from neutralizing cells
into the confinement region.

Table 10.1 Applications of cryogenic technology


Industrial gas production and uses
Gases O2, N2, CO2, He, Ar, Ne, etc
Steel refining (oxygen)
Prestressing of pressure vessels (liquid nitrogen)
Ammonia production (cryogenic systems)
Welding (argon or helium)
Food preservation
Dry ice production
Rapid-freezing with liquid nitrogen
Freeze drying processes
Biomedical applications
Preservation of tissue, blood, semen
Cryosurgery
Mechanical devices
Frictionless bearings and gyroscopes
Refrigeration processes
Electrical devices
Low-noise electronics and masers
Computers
Superconducting motors, generators, energy storage, transmission
Physics studies
Radiation damage
Neutron moderation
Bubble chambers
Low-temperature phenomena, such as superfluidity and superconductivity
Space technology
Rocket propulsion
Space simulation systems and vacuum chambers
Fusion research
Superconducting magnet coils
Cryopanel vacuum pumping systems
10 Cryogenic Systems 493

The cryogenic systems for magnets involve the following considerations:


• Materials. Materials must have suitable low-temperature mechanical, thermal,
and electrical properties.
• Refrigeration. For a reactor heat loads of tens of kW must be removed at
T * 4 K.
• Insulation. Effective insulation must minimize conductive, convective, and
radiative heat transfer.
• Structure. The cryostat must sustain magnetic and gravity forces with a min-
imum of heat leakage.

10.2 Properties of Materials at Low Temperatures

The properties of some materials change markedly at low temperatures. For example,
some organic materials (flowers, banana peels) immersed in liquid nitrogen (at 77 K)
become brittle and will shatter like glass when struck with a hammer.

10.2.1 Mechanical Properties

Some properties of interest are yield stress, ultimate stress, percent elongation,
impact energy, and modulus of elasticity. In general, the percent elongation and
impact energy decrease at low temperatures as shown in Fig. 10.1.

Fig. 10.1 Percent elongation


of various metals as a
function of temperature. Data
from Cryogenic Systems, by
R. Barron, Figs. 2–5, p. 21.
Copyright 1966 used by
permission
494 T. J. Dolan

Fig. 10.2 Ultimate stress of


various materials as a
function of temperature

In the early 1940s, several cargo ships unexpectedly broke apart and sank in the
North Atlantic before it was discovered that their welds became brittle at low
temperatures. Brittle materials usually perform better in compression than in
tension. Because plastics are very brittle at low temperatures, thermal shock must
be avoided. For most materials the yield stress and ultimate stress increase slightly
at low temperatures, as illustrated in Fig. 10.2 for ultimate stress.
Modulus of elasticity usually remains about constant or increases slightly with
decreasing temperature (except for some plastics, where a large increase occurs). If
the magnets are to be repetitively turned on and off, the possibility of low-cycle
fatigue failure should be considered.

10.2.2 Thermal Properties

Pertinent thermal properties include specific heat, thermal conductivity, emissiv-


ity, and thermal expansion coefficient. The specific heats of most inorganic solids
at low temperatures follow the Debye equation
 3 Z hD =T
T dx x4 ex
C ¼ 9R ð10:2Þ
hD 0 ð e x  1Þ 2
where C = specific heat at constant volume (J/mol K), R = 8.314 (J/mol K) is the
universal gas constant, T = material’s temperature (K), and hD is the ‘‘Debye
temperature,’’ which is a constant for a given material. If this equation is divided
by the molecular weight, C will be expressed in J/kg-K. The values of hD for
various materials are shown in Table 10.2.
The ratio of C/R from Eq. (10.2) is plotted as a function of temperature is
shown in Fig. 10.3.
10 Cryogenic Systems 495

Table 10.2 Debye temperatures for Al 390 K


various materials
Cu 310 K
Alpha-Fe 430 K
Gamma-Fe 320 K
Pb 86 K
Li 430 K
Ni 375 K
Ti 350 K

Fig. 10.3 Ratio of specific


heat to universal gas constant
R as a function of the ratio of
temperature T to Debye
temperature hD . From
Cryogenic Systems, by
R. Barron, Figs. 2–8, p. 28.
Copyright 1966, McGraw-
Hill, New York. Used by
permission of McGraw-Hill
Book Company

At temperatures T\hD =12, this equation reduces to

C=R ¼ 233:8 ðT=hD Þ3 ð10:3Þ


The specific heat for stainless steel (18 % Cr, 8 % Ni) falls from 476 J/kg-K at
room temperature to 159 J/kg-K at 77 K and 4.6 J/kg-K at 20 K.
The enthalpy of a material is defined by
Z T
h dT C þ h0 ðJ=kgÞ ð10:4Þ
0

where h0 is the enthalpy at T = 0. The heat added per kg to raise the temperature
from T1 to T2 is
496 T. J. Dolan

Table 10.3 Enthalpies of various materials at low temperatures, J/g


T,K Cu Al Nb Sn Fe Ni C Teflon
300 79.6 170.4 59.2 53.6 81.1 82.4 88.7 167.8
280 72.0 152.5 53.9 49.1 72.3 73.6 75.0 144.6
260 64.4 135.0 48.6 44.7 63.6 65.0 62.5 125.5
240 56.9 117.8 43.4 40.3 55.2 56.7 51.2 107.8
220 49.6 101.0 38.2 36.0 47.0 48.6 41.1 91.3
200 42.4 84.8 33.1 31.7 39.2 40.8 32.2 75.9
180 35.3 69.2 28.0 27.4 31.8 33.4 24.5 61.7
160 28.5 54.4 23.1 23.2 24.6 26.3 18.0 49.0
140 22.1 40.7 18.3 19.1 18.2 19.7 12.7 37.7
120 16.1 28.4 13.8 15.1 12.4 13.8 8.4 27.9
100 10.6 17.8 9.6 11.1 7.6 8.6 5.1 19.5
80 6.0 9.4 5.8 7.6 3.8 4.6 2.7 12.5
60 2.6 3.6 2.8 4.3 1.4 1.8 1.2 7.0
40 0.6 0.8 0.8 1.8 0.3 0.4 0.4 3.0
20 0.03 0.05 0.07 0.3 0.03 0.04 0.04 0.5
10 0.0024 0.005 0.007 0.02 0.005 0.007 0.003 0.05

Z T2
W=M ¼ dT C ¼ h2  h1 ð10:5Þ
T1

where W = heat added (J), M = mass of the material (kg), and h2 and h1 are the
enthalpies at T1 and T2. Enthalpies of several materials are given in Table 10.3.
Example Problem 10.1: Temperature after quench A 1 GJ coil stabilized by
aluminum at 10 K is not to exceed 100 K after a quench in which all the energy is
distributed uniformly in the aluminum, ignoring heat absorption in superconductor
and insulators. What is the required mass of aluminum?
For aluminum, the enthalpy difference h2 - h1 = 17,800 J/kg = W/M.
Since W = 109 J, we find that M = 56,000 kg of aluminum.
Thermal conductivities vs. temperature for various materials are shown in
Fig. 10.4.
Structural materials extending between the coil and the cryostat (container)
should have low thermal conductivity and high strength, so the ratio of yield stress
to thermal conductivity is often used as a figure of merit. The mean value of these
ratios between room temperature and 90 K, relative to the ratio for stainless steel
304, is given in Table 10.4 for various materials.
The linear thermal expansion is used to calculate thermal stresses, which arise
from temperature gradients. Thermal expansion values for various materials are
listed in Table 10.5, relative to the length of the material at 0 K.
The thermal emissivity is the ratio of the amount of heat radiated by a surface to
that which would be radiated by a pure ‘‘black body.’’ Values of thermal emissivity
will be used in estimating radiative heat transfer between walls separated by a vac-
uum. The thermal emissivities of some common materials are given in Table 10.6.
10 Cryogenic Systems 497

Fig. 10.4 Thermal


conductivities of various
materials versus temperature

Table 10.4 Ratio of strength Stainless steel 304 1.00


to thermal conductivity
Aluminum 2024-0 0.017
between room temperature
Teflon 0.269
and 90 K, relative to the ratio
for 304 stainless steel K Monel 0.68
Hastelloy C 0.606
Nylon 0.957
Glass fibers 3.63
Mylar 5.67
Kel F oriented fibers 10.8
Dacron fibers 12.5
From Cryogenic Systems, by R. Barron, Tables 7–5, p. 466.
Copyright 1966, McGraw-Hill, New York. Used by permission
of McGraw-Hill Book Company
498 T. J. Dolan

Table 10.5 Thermal expansion of various materials


Material 20 K 60 K 100 K 150 K 200 K 250 K 300 K
Copper 0 10 44 105 178 256 339
Aluminum 0 10 46 121 214 319 431
Nickel 0 4 23 64 117 172 239
Titanium 0 1 14 41 74 113 155
Stainless 304 -1.1 2.8 30 85 151 225 304
Yellow brass 0 15 57 130 214 303 397
Polyester and fiberglass 3 21 49 99 159 223 291
Nylon 10 81 217 443 716 1,050 1,450
Teflon 45 200 410 717 1,130 1,747 2,695
Data from Scott (1959), Tables 10.5 and 10.6, p. 331 and 333. Some interpolations are given
here. [L(T) - L(0)]/L(0) in units of 10-5

Table 10.6 Thermal emissivities of some common materials


Material T = 77 K T = 293 K
Bright aluminum foil 0.018 0.03
Polished copper 0.019 0.03
Glass 0.94
Stainless steel 304 0.061 0.150
Titanium 0.11
Silver 0.01 0.022
Nickel 0.022 0.04

10.2.3 Electrical Resistivity

Resistivities of Cu and Al, relative to the values at room temperature, are shown in
Fig. 10.5.
These values vary strongly with the impurity content, and the purity is some-
times spoken of in terms of the ratio of the resistivity at 293 K to that at 4.2 K,
called the residual resistivity ratio (RRR), which is about 30 for the aluminum of
Fig. 10.5. Values of RRR of several thousand can be obtained by careful refining,
though at increased cost. The resistivities increase with applied magnetic field and
with neutron irradiation fluence.

10.2.4 Cryogenic Liquids

At cryogenic temperatures ordinary gases become liquids or solids. Properties of


some cryogenic fluids are listed in Table 10.7.
The gases cannot be liquefied above the ‘‘critical temperature,’’ no matter how
high a pressure is applied. Only about 1.3 ppm of ordinary helium is helium-3, the
majority being helium-4. Helium-4 cannot be solidified at atmospheric pressure,
10 Cryogenic Systems 499

Fig. 10.5 Ratio of the


electrical resistivities of Cu
and Al to their resistivities at
293 K. Very pure samples
would have lower
resistivities. From Cryogenic
Systems, by R. Barron,
Figs. 2–12, p. 42. Copyright
1966, McGraw-Hill, New
York. Used by permission of
McGraw-Hill Book Company

Table 10.7 Properties of cryogenic liquids and water at 1 atm


Helium-4 Hydrogen Nitrogen Oxygen Water
Boiling point (K) 4.2 20.3 77.4 90.2 373.2
Melting point (K) None 13.8 63.2 54.4 273.2
Critical temperature (K) 5.2 33.2 126 155 645
Properties at boiling temperature
Density (kg/m3) 125 71 800 1,140 1,000
Heat of vaporization (MJ/m3) 2.72 31.6 161.3 243 2,257
Specific heat cp (kJ/kg-K) 4.56 9.76 2.04 1.70 4.22
Thermal conductivity (W/m-K) 0.0269 0.118 0.139 0.148 0.023
Data from Laquer (1973), Table I, and Cryogenics Systems by R. Barron, Tables 2–3, p. 38.
Copyright 1966 McGraw-Hill, New York. Used by permission of McGraw-Hill Book Company

but it can be solidified at about 25 atm. Liquid helium has many unusual prop-
erties, such as superfluidity, which are described in the references.
Figure 10.6 shows vapor pressures of common fluids at low temperatures.
These curves are taken into account when designing cryogenic pumping sys-
tems. At T = 85 K oxygen liquefies, but nitrogen does not, so hazardous amounts
of liquid oxygen could accumulate.
500 T. J. Dolan

Fig. 10.6 Vapor pressure versus temperature for some common fluids. From H. Neumann, KIT
Summer School, 2010, slide 9

10.3 Refrigeration and Liquefaction

Liquefaction refers to the process of condensing gases into liquids, while refrig-
eration refers to the process of heat removal from a substance. If superconducting
magnets are cooled by liquid helium, then helium liquefaction will be necessary; if
they are cooled by chilled helium vapor, then only refrigeration may be needed.
Both processes are similar, however, in that they require cycles involving com-
pression (with a temperature rise), cooling by a heat sink, and then expansion
cooling (with or without condensation to the liquid phase), as illustrated in the
simple diagram of Fig. 10.7.
According to the Joule-Thomson effect, expansion of a gas lowers its temper-
ature, provided that its initial temperature is below the inversion temperature.
Gases like nitrogen and oxygen, which have inversion temperatures greater than
room temperature, can be cooled by expansion at room temperature. Household
refrigerators use this cycle with a fluid like Freon. (Nowadays, industries cannot
use chlorinated fluorocarbons.) On the other hand, hydrogen and helium have
inversion temperatures of about 204 and 40 K, so they require precooling below
these temperatures if they are to be cooled by expansion. Such gases have been
called ‘‘permanent gases’’ because of the difficulties encountered in attempting to
liquefy them.
In liquefiers, the gas to be liquefied is usually the working fluid. It is precooled
below the inversion temperature, and during expansion part of it condenses into
the liquid state, forming a pool at the bottom of the expansion chamber. To cool
10 Cryogenic Systems 501

Fig. 10.7 A simple


refrigeration cycle. Cryogenic
refrigerators use similar steps,
but with additional measures
to improve efficiency

and liquefy 1 mol of helium starting at 298.15 K, about 27.2 kJ of thermal energy
must be removed. However, a much greater amount of energy must be expended
by the machine to accomplish this amount of heat removal. For liquefaction of
helium, the precooling below the inversion temperature can be done either by
evaporation of liquid hydrogen (which boils at 20.3 K) or by using heat
exchangers with returning chilled helium vapor as the coolant. The latter method
has the advantage of not requiring liquid hydrogen and a separate liquefier for it.
Approximately 280 W of electricity are required per Watt of heat removed at 4 K.
Ordinary expansion through a valve or nozzle is thermodynamically irrevers-
ible: heat energy is wasted in increasing the entropy, and cooling is inefficient.
(Entropy is the ratio of energy to temperature, and it measures the amount of
energy which is bound up in matter and unavailable for use). The expansion can be
made more efficient by letting the gas expand gradually in an expansion engine,
such as a cylinder-and-piston or a gas turbine. In such an engine, the expansion
becomes more nearly reversible or isentropic (constant entropy), and less thermal
energy is trapped in entropy. Consequently, the amount of cooling produced by the
expansion of a given quantity of gas is greatly increased. For this reason expansion
engines are often employed in cryogenic refrigeration and in liquefaction of per-
manent gases.
The Collins helium-liquefaction system is illustrated in Fig. 10.8.
In this system, helium gas at room temperature and atmospheric pressure is
compressed to about 14 atm, and then it is cooled in a series of heat exchangers
with counterflowing chilled helium vapor. Any number of expansion engines may
be used. For simplicity only two are shown in the figure. For this case about 25 %
of the mass flow rate would be channeled through the first expansion engine and
502 T. J. Dolan

Fig. 10.8 The collins system


for helium liquefaction.
Based on Cryogenic Systems
by R. Barron, Figs. 3–30,
p. 122. Copyright 1966,
McGraw-Hill, New York.
Used by permission of
McGraw-Hill Book Company

about 50 % through the second engine, leaving about 25 % to pass through the
Joule-Thomson expansion valve. About 12 % of the mass flow condenses in the
expansion chamber, and the remaining 13 % vapor returns up through the heat
exchangers, where it is joined by returning chilled vapor from the expansion
engines.
In some refrigeration systems regenerators are used in place of heat exchangers.
A regenerator is a heat storage device with a large internal surface area from which
heat may be rapidly transferred to the working fluid. For example, a tube packed
full of steel wool could be used as a regenerator. A regenerator first receives heat
from the hot fluid and then gives up this heat to the cold fluid. The two fluids flow
through it alternately. Regenerators have the advantage of a low pressure drop,
ease of construction, and a very large internal surface area per unit volume. They
are limited by their low heat capacity (which may require rapid interchanging) and
by the mixing of fluids from the two streams as they are interchanged. In practice,
two regenerators may be used simultaneously, one for the hot fluid and one for the
cold, and they may be rapidly interchanged to permit almost continuous flow of the
two fluids.
For simplicity, only the Collins system has been illustrated here. Detailed
analyses of the efficiencies of various types of heat exchangers, regenerators,
expansion engines, and systems for refrigeration and liquefaction are given by
Barron (1966). If high temperature superconducting coils could be used at 65 K,
10 Cryogenic Systems 503

then the cryostat could be much simpler, and the refrigeration power could be
substantially reduced (from 33 to 20 MW in the example of Sect. 4.9).

10.4 Insulation

Effective insulation must combat all three heat transfer mechanisms: conduction,
convection, and radiation. The thermal power flow by conduction along a body
with cross sectional area A and length L is
Z
A T2
Pcond ¼ dT kðTÞ ðWÞ; ð10:6Þ
L T1
where k(T) is the thermal conductivity (W/m-K), and T1 and T2 are the temper-
atures at the endpoints of the body. To minimize the heat flow, one can use
materials with low thermal conductivity, small cross sectional area A, or a long
heat-flow path length L. Some values of the integral are given in Table 10.8.
Example Problem 10.2: Heat leak through support A coil at 10 K is partially
supported by a stainless steel tube with D = 7 mm, thickness 1 mm, and length
20 cm to an intermediate surface at 80 K. What is the heat flow into the coil along
this tube?
From Table 10.8 the difference in conductivity
 integrals is 3.46 W/cm. The
cross sectional area of the tube A ¼ p r22  r21 ¼ 0:251 cm2 . Then from Eq. (10.6)
Pcond ¼ 0:043 W.
Convection is reduced by dividing the gas into small cells (as in foam insula-
tion) or by removing the gas (as in vacuum insulation). When both conduction and
convection have been minimized, radiation may become the dominant heat
transfer process.
The thermal power transferred radiatively from a spherical or cylindrical sur-
face with area A2, temperature T2, and emissivity e2 to another concentric surface
with area A1, temperature T1, and emissivity e1 is

rðT41  T42 Þ
Prad ¼ 1 1
ð10:7Þ
A1 e1 þ A2 e2 ð1  e2 Þ
where r ¼ 5:67  108 W=m-K4 is the Stefan-Boltzmann constant. To minimize
radiant heat transfer, surfaces with low emissivity are used, and multilayer radi-
ation barriers may be employed. For example, if e1 ¼ e2 ¼ 0:90, and if ten radi-
ation shields with emissivity es ¼ 0:05 are placed in between A1 and A2, then the
radiant heat transfer rate Prad drops to only 0.3 % of what it would be without the
shields.
504 T. J. Dolan

RT
Table 10.8 Thermal conductivity integrals (Stewart 1961) 4 dT kðTÞ ðW/mÞ
T, K Electrical tough (Pb) Brass Aluminum 6063-T5 Stainless steels Glass
pitch copper
6 8.00 0.053 0.850 0.0063 0.00211
8 19.1 0.129 2.05 0.0159 0.00443
10 33.2 0.229 3.60 0.0293 0.00681
15 80.2 0.564 9.00 0.0816 0.0131
20 140 1.12 16.5 0.163 0.0200
25 208 1.81 25.8 0.277 0.0279
30 278 2.65 36.5 0.424 0.0368
35 345 3.36 48.8 0.607 0.0471
40 406 4.76 62.0 0.824 0.0586
50 508 7.36 89.5 1.35 0.0846
60 587 10.4 117 1.98 0.115
70 651 13.9 143 2.70 0.151
76 686 16.2 158 3.17 0.175
80 707 17.7 167 3.49 0.194
90 756 22.0 190 4.36 0.240
100 802 26.5 211 5.28 0.292
120 891 36.5 253 7.26 0.408
140 976 47.8 293 9.39 0.542
160 1,060 60.3 333 11.7 0.694
180 1,140 73.8 373 14.1 0.858
200 1,220 88.3 413 16.6 1.03
250 1,420 128 513 23.4 1.50
300 1,620 172 613 30.6 1.99

Although Eq. (10.6) is only valid for simple cases of conductive heat transfer, it
is convenient to define an ‘‘apparent mean thermal conductivity’’ kapp such that
P ¼ kapp A ðT2  T1 Þ=L ð10:8Þ
where P is the total heat flow from all three processes. This definition, although it
is not rigorously accurate, facilitates comparison of kapp for various types of
insulation, as illustrated in Table 10.9.

Table 10.9 Apparent mean thermal conductivities for various types of insula-
tion, between 300 and 77 K. From Glaser et al. (1967) and Vance and Duke
(1962)
Material (lW/cm-K) mW/m-K
Foams, 1 atm 300–400 30–40
Evacuated foams 100–200 10–20
Unevacuated powders 200–1,000 20–100
Evacuated powders (p B 0.01 Pa) 2–20 0.2–2
Evacuated multilayer re-reflective 0.1–2 0.01–0.2
barriers (p B 0.01 Pa)
10 Cryogenic Systems 505

The multiple-barrier insulation, which consists of alternate layers of reflective


aluminum or copper foil and insulating substances, such as thin mesh or plastic or
fiberglass, is clearly the best. Very thin sheets of aluminized Mylar incorporate
both the plastic and the reflector into one sheet. It has been found that kapp rises
rapidly as the pressure is raised above 0.03 Pa. Holes about 5 mm diameter
occupying about 5 % of the sheet surface facilitate vacuum pumping without
spoiling the insulation. As the number of layers of this ‘‘superinsulation’’ is
increased, radiant heat transfer is drastically reduced, but when they are packed too
tightly, conduction becomes significant. The optimum layer density is about 25
layers per cm.
Compaction of the layers must be avoided, as it can increase kapp by an order of
magnitude. One advantage of evacuated powders over superinsulation is ease of
installation, especially around complex shapes but the lower value of kapp provided
by super insulation is probably worth the extra cost of installation. Various types
of multilayer insulation are reviewed by Glaser et al. (1967).
Figure 10.9 shows some ways of installing superinsulation.
The heat leak into liquid helium dewars can also be reduced by using vapor
shielding. The vaporized helium from the inner vessel passes through a tube to
cool an intermediate vessel as it escapes. In this way the temperature difference
seen by the inner vessel is reduced, and boiloff may be reduced by about a factor of

Poor

Better

Best

Fig. 10.9 Methods of installing superinsulation. The overlapping method (top right) is the
poorest, and the individual sheet overlapping (bottom right) is the best, but the most time-
consuming (Neumann 2010)
506 T. J. Dolan

4. High-current leads for superconducting magnets are usually vapor-shielded


tubes, and high-temperature superconductors may be used for part of the lead
where T \ 65 K.

10.5 Cryostats

The vessels used to contain and insulate superconducting magnets are called
cryostats or dewars. The main problem is to sustain the magnetic and gravity
forces without introducing large heat leaks through the structure. Part of the
structure may be at coil temperature (cold reinforcement) and some parts may
make the transition up to 77 and 293 K (warm reinforcement). Cold reinforcement
does not add to heat inflow during normal operation, but steels with good low-
temperature properties must be used, and more metal must be cooled down ini-
tially. If much structural steel is involved, then using plain steel at room tem-
perature may result in cost savings (Powell and Bezler 1974).
A cryostat may have multiple heat shields to reduce heat leaks into the cryo-
genic liquid. Figure 10.10 shows a cryostat for long-term storage of liquid
nitrogen.

Fig. 10.10 A liquid nitrogen vessel with built-in pressure generation (Neumann 2010, slide 14)
10 Cryogenic Systems 507

Fig. 10.11 Supports for a large toroidal field coil. The coil bore is about 7 m wide and 11 m high
(Buncher et al. 1976)

Heat leaks boil some of the LN2, generating pressure, which is measured by the
manometer. Multiple heat shields minimize the heat leak.
The support for a large TF coil is shown in Fig. 10.11.
The bucking cylinder at coil temperature sustains the centering forces, and thin
Kromarc spokes transmit gravity and lateral fault forces out to the dewar (at 77 K).
The dewar is insulated from room temperature by foam plastic. The heat loads
estimated for a torus consisting of 16 such TF coils, producing B = 12 T at the
coils, are listed in Table 10.10.
The losses in the pump are due to heat added in compressing the liquid helium
coolant and overcoming frictional pressure drops.
Cryostats need to have instrumentation to measure temperature, pressure, mass
flow rate, cryogenic fluid level, and stress.

10.6 ITER Cryogenic System

The ITER cryogenic system will provide an average cooling power of 65 kW at


4.5 K and a peak power of 1.3 MW at 80 K. The liquid helium system must be
able to accommodate the large pulsed heat loads from the magnets. It must be able
to support plasma pulses lasting up to 50 min at hundreds of MW, including the
pulsed loads of the central solenoid and regeneration of the NBI cryopumps (Sect.
9.7).
Figure 10.12 shows the ITER cryogenic system.
The magnetic coils and cryopumps are cooled by forced supercritical flow of
liquid helium (SHe) from distribution boxes at 3.7–4.2 K using cold circulators
508 T. J. Dolan

Table 10.10 Helium Structural conduction kW


refrigeration requirements for
an array of 16 TF coils Vertical spokes 0.075
producing a peak field of 12 T Horizontal spokes 2.941
at the coils (Buncher et al. Residual gas conduction
1976) Dewar 0.472
Vertical spokes 0.036
Horizontal spokes 0.732
Thermal radiation
Dewar 0.321
Vertical spokes 0.017
Horizontal spokes 0.355
Nuclear radiation
Conductor 0.156
Structure 0.260
AC losses
Conductor 12.9
Structure 3.2
Joint losses 17.3
Losses in pump
(*50 % efficiency pump) 29.2
Total thermal load Pth 68.0
Electrical load Pe & 280 Pth 19000

Fig. 10.12 A simplified


diagram of a supercritical
helium system. The helium
bath removes heat from the
SHe loop, and the LHe bath is
replenished by the cryoplant
(Courtesy of ITER
Organization)
10 Cryogenic Systems 509

and long length transfer lines within the Tokamak building. Heat exchangers at the
inlet and outlet of the cold circulators are employed to transfer the heat loads to the
helium bath while maintaining stable operating conditions. Liquid helium will be
stored in a 120 m3 dewar. Three liquid helium refrigerators operate in parallel to
supply and recover helium from the main users and provide the required cooling
power at 4 and 50 K. An air separation unit will produce liquid nitrogen (LN2)
on-site for the liquid helium refrigerators, the 80 K gaseous helium loop and the
auxiliary systems. The 80 K helium loop provides cooling for the machine and
components thermal shields (Serio 2012).
Liquid helium (LHe) is stored in 3 large tanks (baths), which remove heat via a
heat exchanger (Hx) from the supercritical helium (SHe) that flows through the
magnet coils and cryopumps, Fig. 10.12.
The cryodistribution system carries liquid helium and nitrogen from the cryo-
plant building to the tokamak building via the plant bridge. The cryogenic dis-
tribution lines have a total length of 3.7 km. The Tokamak building and equipment
(including cryogenic components) must be protected against earthquakes by iso-
lation pads. The system includes 7 large cryogenic distribution boxes, a cold
compressor box, a large interconnecting box, 31 magnet feeders housing the
valves, instrumentation, and flow control devices to maintain safe, reliable oper-
ation during cool-down, warm-up and plasma operations.
During a 16 months experimental campaign, the cryogenic system will operate
24 h per day for 11 consecutive days, followed by 2–3 days maintenance. More
than 200 possible cryogenic system failure modes and 500 safety hazard failure
modes have been studied, and mitigation actions have been implemented. The
overall cryogenic system availability is expected to be above 97 %, and no major
critical safety failure mode remains (Serio 2012).
In conclusion, cryogenic engineering is a mature discipline, and the large,
complex ITER cryogenic systems should be able to operate safely and reliably.

10.7 Problems

10.1. Assume that the fiberglass support tubes of Fig. 10.13 have outer radius
28 cm, inner radius 27 cm and length 28 cm, and that the superinsulation is
1 cm thick with kapp ¼ 104 W=m-K. The copper leads have a total cross
sectional area of 0.1 cm2 and lengths of 0.9 m. Assume that the thermal
conductivity integral of the fiberglass is about twice that of glass, and that
the coils may be considered to be all copper in estimating their mass and
enthalpy. Estimate (a) the coil mass (read dimensions from sketch), (b) heat
flow rate along the fiberglass support tubes, (c) heat flow rate along the coil
leads, (d) heat leak rate through the superinsulation.
10.2. Estimate the heat which must be removed and the number of liters of liquid
nitrogen that are boiled off in cooling 1 kg of copper from 300 to 80 K, and
510 T. J. Dolan

Fig. 10.13 The cryostat for a pair of spindle cusp magnet coils. The helium fill lines and coil
current leads enter vertically from above (not shown)

the how many liters of liquid helium are boiled off in cooling it from 80 to
4.2 K. How many liters of nitrogen and helium will be consumed in cooling
down the coils of problem 1, ignoring other losses?
10.3. How many liters of liquid helium will be boiled off per hour by a heat input
of 1 kW? By the total heat input to the coils of Problem 1?

10.8 Review Questions

1. What invention by Collins greatly facilitated development of low-temperature


research?
2. What are some applications of cryogenic technology?
RT2
3. Explain the equation W=M ¼ dT C ¼ h2  h1
T1
4. For what materials is the ratio of strength/(thermal conductivity) important?
10 Cryogenic Systems 511

5. Define ‘‘critical temperature’’ and ‘‘inversion temperature’’.


6. Sketch a simple refrigeration cycle and explain its operation.
7. What is the advantage of using an expansion engine instead of a simple nozzle
for cooling of a gas?
8. Explain how a Collins liquefier works.
RT2
9. Explain the equation P ¼ ðA=LÞ dT kðTÞ
T1
10. Explain the meaning of the equation P ¼ kapp A ðT2  T1 Þ=L
11. What is vapor shielding, and where is it useful?
12. What instrumentation is needed for a cryostat?

References

Barron R (1966) Cryogenic systems. McGraw-Hill, New York


Buncher BR, Chi JWH, Fernandez R (1976) Conceptual studies of toroidal field magnets for the
tokamak experimental power reactor. Final report, ORO-5153-1. Westinghouse Electric
Corporation, Pittsburgh
Glaser PE, Black IA, Lindstrom RS, Ruccia FE, Wechsler AE (1967) Thermal insulation systems,
National Aeronautics and Space Administration report NASA SP-5027. Washington, DC
Laquer HL (1973) Cryogenics, the uncommon cold. USAEC Technical Information Center, Oak
Ridge
Neumann H (2010) Cryogenics. Summer School on Fusion Technology, Karlsruhe Institute of
Technology, Germany
Powell JR, Bezler P (1974) Warm reinforcement and cold reinforcement magnet systems for
tokamak fusion power reactors: a comparison, technology of controlled thermonuclear fusion
experiments and the engineering aspects of fusion reactors, CONF-721111, USAEC, pp 358.
Scott RB (1959) Cryogenic engineering. Van Nostrand, Princeton
Serio L (2012) Challenges for cryogenics in the nuclear fusion quest: the ITER cryogenic system.
In: 24th international cryogenic engineering conference, Fukuoka, 14–18 May 2012
Stewart RB, Johnson VJ (1961) A compendium of properties of materials at low temperatures,
WADD technical report 60-56, part IV
Vance RW, Duke WM (eds) (1962) Applied cryogenic engineering. Wiley, New York

You might also like