0% found this document useful (0 votes)
454 views895 pages

Cryogenic Engineering

This document provides an overview of the second edition of the book "Cryogenic Engineering" by Thomas M. Flynn. Some key details: - It is a revised and expanded second edition of the book intended for professional engineers and physicists working with cryogenics. - The book aims to provide practical knowledge, data, examples, and guidelines that can be directly applied to cryogenic engineering problems and systems design. - It builds upon decades of cryogenic engineering research and technology developed at the National Bureau of Standards Cryogenic Engineering Laboratory under the leadership of Russell Scott and others.

Uploaded by

kaizhang0222
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
0% found this document useful (0 votes)
454 views895 pages

Cryogenic Engineering

This document provides an overview of the second edition of the book "Cryogenic Engineering" by Thomas M. Flynn. Some key details: - It is a revised and expanded second edition of the book intended for professional engineers and physicists working with cryogenics. - The book aims to provide practical knowledge, data, examples, and guidelines that can be directly applied to cryogenic engineering problems and systems design. - It builds upon decades of cryogenic engineering research and technology developed at the National Bureau of Standards Cryogenic Engineering Laboratory under the leadership of Russell Scott and others.

Uploaded by

kaizhang0222
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd

CRYOGENIC

ENGINEERING
Second Edition
Revised and Expanded

© 2005 by Marcel Dekker


CRYOGENIC
ENGINEERING
Second Edition
Revised and Expanded

Thomas M. Flynn
CRYOCO, Inc.
Louisville, Colorado, U.S.A.

M ARCEL D EKKER N EW Y ORK

© 2005 by Marcel Dekker


Although great care has been taken to provide accurate and current information, neither the
author(s) nor the publisher, nor anyone else associated with this publication, shall be liable for
any loss, damage, or liability directly or indirectly caused or alleged to be caused by this book.
The material contained herein is not intended to provide specific advice or recommendations
for any specific situation.

Trademark notice: Product or corporate names may be trademarks or registered trademarks


and are used only for identification and explanation without intent to infringe.

Library of Congress Cataloging-in-Publication Data


A catalog record for this book is available from the Library of Congress.

ISBN: 0-8247-5367-4

This book is printed on acid-free paper.

Headquarters
Marcel Dekker, 270 Madison Avenue, New York, NY 10016, U.S.A.
tel: 212-696-9000; fax: 212-685-4540

Distribution and Customer Service


Marcel Dekker, Cimarron Road, Monticello, New York 12701, U.S.A.
tel: 800-228-1160; fax: 845-796-1772

World Wide Web


http:==www.dekker.com

The publisher offers discounts on this book when ordered in bulk quantities. For more infor-
mation, write to Special Sales=Professional Marketing at the headquarters address above.

Copyright # 2005 by Marcel Dekker, All Rights Reserved.


Neither this book nor any part may be reproduced or transmitted in any form or by any
means, electronic or mechanical, including photocopying, microfilming, and recording, or
by any information storage and retrieval system, without permission in writing from the
publisher.

Current printing (last digit):

10 9 8 7 6 5 4 3 2 1

PRINTED IN THE UNITED STATES OF AMERICA

© 2005 by Marcel Dekker


The places I took him!
I tried hard to tell
Young Conrad Cornelius o’Donald o’Dell
A few brand new wonderful words he might spell.
I led him around and I tried hard to show
There are things beyond Z that most people don’t know.
I took him past Zebra. As far as I could.
And I think, perhaps, maybe I did him some good . . .
On Beyond Zebra, by Dr. Seuss.
With permission, Random House Inc. 1745 Broadway, New York, NY.

To Rita

© 2005 by Marcel Dekker


Preface to the Second Edtion

Dr. Flynn’s Cryogenic Engineering 2nd Edition was written for a specific audience,
namely, the professional engineer or physicist who needs to know some cryogenics
to get his or her job done, but not necessarily make a career of it. The 2nd Edition
was written to follow closely the cryogenic engineering professional course given
annually by Dr. Flynn for 25 years, and accordingly has been thoroughly tested
to be very practical. This 2nd (and last) edition includes over 125 new literature
citations, and features more than 130 new graphs and tables of data, which may
no longer be available elsewhere.

© 2005 by Marcel Dekker


Preface to the First Edition

This book is for the engineer and scientist who work for a living and who need cryo-
genics to get a job done. It is a deskbook, containing hundreds of tables and chart of
cryogenic data that are very hard to come by. Examples and sample calculations of
how to use the data are included. It is not a textbook. Instead, it assumes that the
reader already has basic engineering and science skills. It is practical, using the mea-
surement units of trade—SI, U.S. customary, and hybrid systems—just as they are
commonly used in practice. It is not a design text, but it does contain many useful
design guidelines for selecting the right system, either through procurement or
in-house construction. In short, it is the cryogenics book I would like to have on
my desk.
This book was written to gather into one source much of the technology devel-
oped at the National Bureau of Standards (NBS) Cryogenic Engineering Laboratory
in Boulder, COL, over the last 40 years.
In the early 1950s, there was a need for the rapid development of a liquid
hydrogen technology, and the major responsibility for the progress of this new engi-
neering specialty was entrusted to the Cryogenics Section of the Heat and Power
Division of NBS in Washington, D.C. Russell Scott led the work as chief of that sec-
tion. Scott soon became the individual immediately in charge of the design and con-
struction in Boulder (in March 1952) of the first large-scale liquefier for hydrogen
ever built. This was the beginning of the Cryogenic Engineering Laboratory of NBS.
Again, in the late 1950s—when the nation was striving to regain world leader-
ship in the exploration of space—the NBS Cryogenics Engineering Laboratory,
which had by then matured under Scott’s leadership, assumed a pre-eminent role
in the solution of problems important to the nation. Scott, having had the foresight
to establish a Cryogenic Data Center within the laboratory, was able to provide a
focal point for information on many aspects of cryogenic engineering.
As a result of all this pioneering in the field of low-temperature engineering, a
considerable amount of valuable technology was developed that in the course of nor-
mal events might have been lost. Scott recognized this, and the result was his book
Cryogenic Engineering, an important first in its field. Its quality, authority, and com-
pleteness constitute a lasting tribute to him.
This present book is a mere shadow of Scott’s work but is intended once more
to up-date and preserve some of the cryogenics developed at NBS. There is only one
author’s name on the cover. Nonetheless, this book is a product of the collaboration
of hundreds of good men and women of the NBS Cryogenic Engineering Labora-
tory. It is written to preserve the technology they developed.

vii

© 2005 by Marcel Dekker


viii Preface

When I was about to graduate as a chemical engineer from Rice University in


1955, I proudly told my department chairman, Dr. Arthur J. (‘‘Pappy’’) Hartsook,
that I intended to go to graduate school. Pappy, who knew I was a mediocre student,
brightened considerably when I told him I wasn’t going to a ‘‘good’’ school, like
MIT, Michigan, or Wisconsin. Instead, I was going to the University of Colorado,
where I could learn to ski. Dr. Hartsook was so relieved that he gave me a piece of
advice pivotal to my career and my life. I will share it with you now. Pappy said that
What I would work on was not nearly as important as who I would work for.
I took that advice and chose to work for a new professor at the University of
Colorado, Dr. Klaus Timmerhaus. Klaus had only been there a year or two; the
National Science Foundation and grantsmanship had yet to be invented. I had a
teaching assistantship (paper grader) at $150 per month, before taxes. It was the
most money I had ever had. To help me get some money for our planned research,
Klaus suggested that I work at the Cryogenic Engineering Laboratory of the
National Bureau of Standards. And so I did, for the next 28 years. For many years,
it was the best of times, a truly nurturing environment for the young engineer scien-
tist, because of the people who either worked there or visited there.
I wish to thank Klaus Timmerhaus, Russell Scott, Bascom Birmingham, Dud-
ley Chelton, Bob Powell, John Dean, Ray Smith, Jo Mandenhall, Jim Draper, Dick
Bjorklund, Bill Bulla, M. D. Bunch, Bob Goodwin, Lloyd Weber, Ray Radebaugh,
Peter Storch, Larry Sparks, Bob McCarty, Vic Johnson, Bill Little, Bob Paugh, Bob
Jacobs, Mike McClintock, Al Schmidt, Pete Vander Arend, Dan Weitzel, Wally
Ziegler, John Gardner, Bob Mohling, Bob Neff, Scott Willen, Will Gully, Art Kid-
nay, Graham Walker, Ralph Surloch, Albert Schuler, Sam Collins, Bill Gifford,
Peter Gifford, Ralph Longsworth, Ed Hammel, and Fred Edeskuty. Special thanks
are due to Chris Davis and Janet Diaz for manuscript preparation and technical edit-
ing. I mention all these names not so much to give credit as to spread the blame.
I apologize in advance to those I have forgotten to mention.

© 2005 by Marcel Dekker


Contents

Preface to the Second Edition . . . . v


Preface to the First Edition . . . . vii

1. Cryogenic Engineering Connections . . . . . . . . . . . . . . . . . . . . . . . . . 1


1. Forewarned . . . . 1
2. The Entrepreneurs . . . . 1
3. The Butchers . . . . 4
4. The Brewers . . . . 5
5. The Industrialists . . . . 6
6. The Scientists . . . . 9
7. The Engineers . . . . 12
8. The Rocket Scientists . . . . 15
9. The Physicists and Superconductivity . . . . 16
10. Science Marches On . . . . 21

2. Basic Principles . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23
1. Introduction . . . . 23
2. Thermodynamics . . . . 27
3. Heat Transfer . . . . 55
4. Momentum Transfer . . . . 68
5. Heat Leak and Pressure Drop in Cryogenic Transfer Lines . . . . 69
6. Cooldown . . . . 73
7. Summary . . . . 76

3. Cryogenic Fluids . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 77
1. PVT Behavior of a Pure Substance . . . . 77
2. Temperature–Enthalpy and Temperature–Entropy Diagrams of Pure
Substances . . . . 81
3. Properties and Uses of Cryogenic Fluids . . . . 83

4. Mechanical Properties of Solids . . . . . . . . . . . . . . . . . . . . . . . . . 257


1. Introduction . . . . 257
2. Strength, Ductility, and Elastic Modulus . . . . 257
3. The Structure of Solids . . . . 259
4. Ductility . . . . 261

ix

© 2005 by Marcel Dekker


x Preface

5. Low-Temperature Strength of Solids . . . . 268


6. Elastic Constants . . . . 274
7. Modulus of Elasticity . . . . 279
8. Fatigue Strength . . . . 282
9. Mechanical Properties Summary . . . . 284
10. Design Considerations . . . . 285
11. Material Selection Criteria for Cryogenic Tanks . . . . 291

5. Transport Properties of Solids . . . . . . . . . . . . . . . . . . . . . . . . . . 301


1. Thermal Properties . . . . 301
2. Emissivity, Absorptivity, and Reflectivity . . . . 327
3. Electrical Properties . . . . 332
4. Superconductivity . . . . 339

6. Refrigeration and Liquefaction . . . . . . . . . . . . . . . . . . . . . . . . . . 359


1. Introduction . . . . 359
2. Refrigeration and Liquefaction . . . . 359
3. Recuperative Cycles . . . . 367
4. Liquefaction of Gases . . . . 369
5. Refrigerator Efficiency . . . . 377
6. Useful Thermodynamic Relations . . . . 380
7. Refrigeration and Liquefaction Methods . . . . 381
8. Large Systems . . . . 401
9. Regenerative Cycles . . . . 401
10. Magnetocaloric Refrigeration . . . . 427
11. Ultra Low-Temperature Refrigerators . . . . 433
12. Very Small Coolers . . . . 438
13. Superconductors and Their Cooling Requirements . . . . 439
14. Cryocoolers . . . . 441
15. Conclusion . . . . 442

7. Insulation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 445
1. Introduction . . . . 445
2. Heat Transfer . . . . 446
3. Vacuum Insulation . . . . 448
4. Evacuated Porous Insulation . . . . 468
5. Gas-Filled Powders and Fibrous Materials . . . . 475
6. Solid Foams . . . . 477
7. Multilayer Insulation . . . . 483
8. Vapor Barriers . . . . 510
9. Protective Enclosures . . . . 513
10. Liquid and Vapor Shields . . . . 514
11. Composite Insulations . . . . 518
12. Other Materials . . . . 525
13. Placement of Insulation Systems . . . . 525
14. Adhesives . . . . 526

© 2005 by Marcel Dekker


Preface xi

15. Comparison of Insulations . . . . 528


16. Summary . . . . 535

8. Cryogenic Instrumentation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 537


1. Introduction . . . . 537
2. Strain . . . . 539
3. Displacement and Position . . . . 543
4. Pressure . . . . 545
5. Flow . . . . 554
6. Liquid Level . . . . 579
7. Density . . . . 587
8. Temperature . . . . 594

9. Cryogenic Equipment and Cryogenic Systems Analysis . . . . . . . . . 633


1. Introduction . . . . 633
2. Compressors . . . . 638
3. Pumps . . . . 644
4. Expansion Engines . . . . 652
5. Valves . . . . 665
6. Heat Exchangers . . . . 671
7. Storage . . . . 696
8. Transfer of Liquefied Gases . . . . 714

10. Natural Gas Processing and Liquefied Natural Gas . . . . . . . . . . . 727


1. Introduction . . . . 727
2. Purification . . . . 730
3. Hydrocarbon Recovery . . . . 742
4. Cryogenic Upgrading of Natural Gas . . . . 745
5. Helium Extraction, Nitrogen Rejection, and Hydrocarbon
Recovery . . . . 748
6. Liquefaction of Natural Gas . . . . 748

11. Safety with Cryogenic Systems . . . . . . . . . . . . . . . . . . . . . . . . . 773


1. Introduction . . . . 773
2. Physiological Hazards . . . . 773
3. Suitability of Materials and Construction Techniques . . . . 777
4. Explosions and Flammability . . . . 788
5. Excessive Pressure Gas . . . . 806
6. Special Considerations for Hydrogen . . . . 811
7. Special Considerations for Oxygen . . . . 837
8. General Safety Principles . . . . 865
9. Safety Checklist . . . . 869

References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 873

© 2005 by Marcel Dekker


5367-4 Flynn Ch01 R2 090704

1
Cryogenic Engineering Connections

1. FOREWARNED

In his masterpiece, Civilization, Sir Kenneth Clark writes not a foreword but rather a
fore-warned. Clark writes to the effect that he does not know if his recounting is the
actual history of civilization or not. He only knows that his story is his own view of
the history of civilization. Having said that, Clark found much relief from the
tedium of accuracy and was able to tell a more entertaining, and still highly accurate,
story.
I wish to do the same. I do not know if the recounting to follow in this chapter
is accurate or not. I only know that it is my own view of how it happened. I hope it is
accurate—I did not deliberately make any of it up—but who knows?
I am an American and unabashedly proud of that fact. Therefore, let me begin
with the American who may have gotten it all going, John Gorrie (1803–1855).

2. THE ENTREPRENEURS

John Gorrie was born in Charleston South Carolina in 1803 and graduated from
the College of Physicians and Surgeons in New York City in 1833. In the spring of
1833 Gorrie moved to Apalachicola, a small coastal town in Florida situated on
the Gulf of Mexico at the mouth of the river after which the town is named. By
the time he arrived it was already a thriving cotton port, where ships from the
north-east arrived to unload cargoes of supplies and load up with cotton for the
northern factories. Within a year Gorrie was involved with town affairs. He served
as mayor, city treasurer, council member, bank director, and founder of Trinity
Church.
In 1834, he was made postmaster and in 1836 president of the local branch of
the Pensacola Bank. In the same year, the Apalachicola Company asked him to
report on the effects of the climate on the population, with a view to possible expan-
sion of the town. Gorrie recommended drainage of the marshy, low-lying areas that
surrounded the town on the grounds that these places gave off a miasma com-
pounded by heat, damp and rotting vegetation which, according to the Spallanzani
theory with which every doctor was intimate, carried disease. He suggested that only
brick buildings be erected. In 1837, the area enjoyed a cotton boom, and the town
population rose to 1500. Cotton bales lined the streets, and in four months 148 ships
arrived to unload bricks from Baltimore, granite from Massachusetts, house framing
from New York. Gorrie saw that the town was likely to grow as commerce increased
1

© 2005 by Marcel Dekker


5367-4 Flynn Ch01 R2 090704

2 Chapter 1

and suggested that there was a need for a hospital. There was already a small medical
unit in operation under the auspices of the US Government, and Gorrie was
employed there on a part-time basis. Most of his patients were sailors and waterside
workers, and most of them had fever, which was endemic in Apalachicola every
summer.
Gorrie became obsessed with finding a cure to the disease. As early as 1836, he
came close to the answer, over 60 years ahead of the rest of the world. In that year he
wrote: ‘‘Gauze curtains, though chiefly used to prevent annoyance and suffering
from mosquitoes, are thought also to be sifters of the atmosphere and interceptors
and decomposers of malaria.’’ The suggestion that the mosquito was the disease
carrier was not to be made until 1881, many years after Gorrie’s death, and for
the moment he presumed that it came in some form of volatile oil, rising from
the swamps and marshes.
By 1838, Gorrie had noticed that malaria seemed to be connected with hot,
humid weather, and he set about finding ways to lower the temperature of his
patients in summer. He began by hanging bowls full of ice in the wards and circu-
lating the cool air above them by means of a fan. The trouble was that in Apala-
chicola ice was hard to come by. Ever since a Massachusetts merchant named
Frederic Tudor had hit on the idea of cutting ice from ponds and rivers in winter
and storing it in thick-walled warehouses for export to hot countries, regular ice
shipments had left the port of Boston for destinations as far away as Calcutta.
But Apalachicola was only a small port which the ships often missed altogether:
if the ice crop was poor, the price rose to the exorbitant rate of $1.25 a pound.
In 1844, Gorrie found the answer to the problem. It was well known that com-
pressed gases which are rapidly allowed to expand absorb heat from their surround-
ings, so Gorrie constructed a steam engine to drive a piston back and forward in a
cylinder. His machine compressed air, causing it to heat, and then the air through
radiant coils where it decompressed and cooled, absorbing heat from a bath of
brine. On the next cycle the air remained cool, since the brine had given up most
of its heat. This air was then pumped out of the cylinder and allowed to circulate
in the ward. Gorrie had invented air-conditioning. By bringing the cold brine into
contact with water, Gorrie was then able to draw heat from the water to a point
where it froze. Gorrie’s application of compressed and decompressed gas as a cool-
ant in radiant coils remains the common method for cooling air in modern refrig-
eration systems. His first public announcement of this development was made on 14
July 1850 in the Mansion House Hotel, where M. Rosan, the French Consul in
Apalachicola, was celebrating Bastille Day with champagne. No ice ship had
arrived, so the champagne was to be served warm. At the moment of the toast
to the French Republic four servants entered, each carrying a silver tray on which
was a block of ice the size of a house-brick, to chill the wine, as one guest put it,
‘‘by American genius’’.
In May of the following year Gorrie obtained a patent for the first ice-making
machine (see Fig. 1.1), the first patent ever issued for a refrigeration machine. The
patent specified that the water container should be placed in the cylinder, for faster
freezing. Gorrie was convinced his idea would be a success. The New York Times
thought differently: ‘‘There is a crank,’’ it said, ‘‘down in Apalachicola, Florida,
who claims that he can make ice as good as God Almighty!’’ In spite of this, Gorrie
advertised his invention as ‘‘the first commercial machine to work for ice making and
refrigeration.’’ He must have aroused some interest, for later he was in New Orleans,
selling the idea that ‘‘a ton of ice can be made on any part of the Earth for less than

© 2005 by Marcel Dekker


5367-4 Flynn Ch01 R2 090704

Cryogenic Engineering Connections 3

Figure 1.1 Improved process for the artificial production of ice.

$2.00.’’ But he was unable to find adequate backing, and in 1855 he died, a broken
and dispirited man. A statue of John Corrie now stands in Statuary Hall of the US
Capitol building, a tribute from the State of Florida to his genius and his importance
to the welfare of mankind.
Three years after his death a Frenchman, Ferdinand Carré, produced a
compression ice-making system and claimed it for his own, to the world’s acclaim.
Carré was a close friend of M. Rosan, whose champagne had been chilled by
Gorrie’s machine eight years before.
Just before he died, Gorrie wrote an article in which he said: ‘‘The system is
equally applicable to ships as well as buildings . . . and might be instrumental in pre-
serving organic matter an indefinite period of time.’’ The words were prophetic,
because 12 years later Dr. Henry P. Howard, a native of San Antonio, used the
air-chilling system aboard the steamship Agnes to transport a consignment of frozen
beef from Indianola, Texas, along the Gulf of Mexico to the very city where Gorrie
had tried and failed to get financial backing for his idea. On the morning of Saturday
10 June 1869, the Agnes arrived in New Orleans with her frozen cargo. There it was
served in hospitals and at celebratory banquets in hotels and restaurants. The New
Orleans Times Picayune wrote: ‘‘[The apparatus] virtually annihilates space and
laughs at the lapse of time; for the Boston merchant may have a fresh juicy beefsteak
from the rich pastures of Texas for dinner, and for dessert feast on the delicate,
luscious but perishable fruits of the Indies.’’
At the same time that Howard was putting his cooling equipment into the
Agnes, committees in England were advising the government that mass starvation
was likely in Britain because for the first time the country could no longer feed itself.
Between 1860 and 1870, consumption of food increased by a staggering 25%. As the
population went on rising, desperate speeches were made about the end of
democracy and nationwide anarchy if the Australians did not begin

© 2005 by Marcel Dekker


5367-4 Flynn Ch01 R2 090704

4 Chapter 1

immediately to find a way of sending their sheep in the form of meat instead of
tallow and wool.
Thomas Malthus (1766–1834), the first demographer, published in An Essay
on the Principle of Population (1798). According to Malthus, population tends to
increase faster than the supply of food available for its needs. Whenever a relative
gain occurs in food production over population growth, a higher rate of population
increase is stimulated; on the other hand, if population grows too much faster than
food production, the growth is checked by famine, disease, and war. Malthus’s the-
ory contradicted the optimistic belief prevailing in the early 19th century, that a
society’s fertility would lead to economic progress. Malthus’s theory won supporters
and was often used as an argument against efforts to better the condition of the
poor. (the poor should die quietly). Those (the vast majority) who did not read
the complete essay, assumed that mass starvation and imminent and inevitable, lead-
ing to the so-called Malthusian revolution in England. There were food riots in the
streets. Food storage and transportation was seen as a critical global issue facing
mankind.
Partially for these reasons, two Britons, Thomas Mort and James Harrison,
emigrated to Australia and set up systems to refrigerate meat. In 1873 Harrison
gave a public banquet of meat that had been frozen by his ice factory, to cele-
brate the departure of the S.S. Norfolk for England. On board were 20 tons of
mutton and beef kept cold by a mixture of ice and salt. On the way, the system
developed a leak and the cargo was ruined. Harrison left the freezing business.
Mort then tried a different system, using ammonia as the coolant. He too gave
a frozen meat lunch, in 1875, to mark the departure for England of the S.S.
Northam. Another leak ruined this second cargo, and Mort retired from the busi-
ness. But both men had left behind them working refrigeration plants in Austra-
lia. The only problem was to find the right system to survive the long voyage to
London. Eventually, the shippers went back to Gorrie’s ‘‘dry air’’ system. Aboard
ship, it was much easier to replace leaking air than it was to replace leaking
ammonia. Even though ammonia was ‘‘more efficient’’, it was sadly lacking while
air was ubiquitous. This NH3=air substitution is one early example of a lack of
‘‘systems engineering’’. There are often unintended consequences of technology,
and the working together of the system as a whole must be considered. NH3
refrigerators were thermodynamically more efficient, but could not be relished
with NH3 readily.

3. THE BUTCHERS

The development of the air-cycle refrigerator, patented by the Scottish butchers


Bell and Coleman in 1877, made the technical breakthrough. The use of atmo-
spheric air as a refrigerating fluid provided a simple, though inefficient, answer
to ship-board refrigeration and led to the British domination of the frozen meat
trade thereafter.
The first meat cargo to be chilled in this fashion left Australia aboard the S.S.
Strathleven on 6 December 1879 to dock in London on 2 February of the following
year with her cargo intact (Fig. 1.2). Figure 1.3 shows a similar ship to the S.S.
Strathleven being unloaded. The meat sold at Smithfield market for between 5d
and 6d per pound and was an instant success. Queen Victoria, presented with a
leg of lamb from the same consignment, pronounced it excellent. England was saved.

© 2005 by Marcel Dekker


5367-4 Flynn Ch01 R2 090704

Cryogenic Engineering Connections 5

Figure 1.2 The S. S. Strathleven, carrying the first successful consignment of chilled beef
from Australia to England. Note the cautious mixture of steam and sail, which was to
continue into the 20th century.

4. THE BREWERS

Harrison’s first attempts at refrigeration in Australia had been in a brewery, where


he had been trying to chill beer, and although this operation was a moderate success,
the profits to be made from cool beer were overshadowed by the immense potential
of the frozen meat market. The new refrigeration techniques were to become a boon
to German brewers, but in Britain, where people drank their beer almost at room
temperature, there was no interest in chilling it. The reason British beer-drinkers take
their beer ‘‘warm’’ goes back to the methods used to make the beer. In Britain it is

Figure 1.3 Unloading frozen meat from Sydney, Australia, at the South West India Dock,
London. This shows the hold of the Catania, which left port in August 1881 with 120 tons of
meat from the same exporters who had filled the Strathleven.

© 2005 by Marcel Dekker


5367-4 Flynn Ch01 R2 090704

6 Chapter 1

produced by a method using a yeast which ferments on the surface of the beer vat
over a period of 5–7 days, when the ideal ambient temperature range is from 60 F
to 70 F. Beer brewed in this way suffers less from temperature changes while it is
being stored, and besides, Britain rarely experiences wide fluctuations in summer
temperatures. But in Germany beer is produced by a yeast which ferments on the
bottom of the vat. This type of fermentation may have been introduced by monks
in Bavaria as early as 1420 and initially was an activity limited to the winter months,
since bottom fermentation takes place over a period of up to 12 weeks, in an ideal
ambient temperature of just above freezing point. During this time the beer was
stored in cold cellars, and from this practice came the name of the beer: lager, from
the German verb lagern (to store). From the beginning there had been legislation in
Germany to prevent the brewing of beer in the summer months, since the higher tem-
peratures were likely to cause the production of bad beer. By the middle of the 19th
century every medium-sized Bavarian brewery was using steam power, and when the
use of the piston to compress gas and cool it became generally known, the president
of the German Brewer’s Union, Gabriel Sedlmayr of the Munich Spätenbrau brew-
ery, asked a friend of his called Carl Von Linde if he could develop a refrigerating
system to keep the beer cool enough to permit brewing all the year round. Von Linde
solved Sedlmayr’s problem and gave the world affordable mechanical refrigeration,
an invention that today is found in almost every kitchen.
Von Linde used ammonia instead of air as his coolant, because ammonia lique-
fied under pressure, and when the pressure was released it returned to gaseous form,
and in so doing drew heat from its surroundings. In order to compress and release
the ammonia, he used Gorrie’s system of a piston in a cylinder. Von Linde did
not invent the ammonia refrigeration system, but he was the first to make it work.
In 1879 he set up laboratories in Wiesbaden to continue research and to convert
his industrial refrigeration unit into one for the domestic market. By 1891, he had
put 12,000 domestic refrigerators into German and American homes. The modern
fridge uses essentially the same system as the one with which Von Linde chilled
the Spätenbrau cellars.

5. THE INDUSTRIALISTS

Interest in refrigeration spread to other industries. The use of limelight, for instance,
demanded large amounts of oxygen, which could be more easily handled and trans-
ported in liquid form. The new Bessemer steel-making process used oxygen. It may
be no coincidence that an ironmaster was involved in the first successful attempt to
liquefy the gas. His name was Louis Paul Cailletet, and together with a Swiss
engineer, Raoul Pictet, he produced a small amount of liquid oxygen in 1877.
At the meeting of the Académie des Sciences in Paris on 24 December 1887,
two announcements were made which may be recognized as the origins of cryogenics
as we know it today. The secretary to the Académie spoke of two communications he
had received from M. Cailletet working in Paris and from Professor Pictet in Geneva
in which both claimed the liquefaction of oxygen, one of the permanent gases.
The term ‘‘permanent’’ had arisen from the experimentally determined fact
that such gases could not be liquefied by pressure alone at ambient temperature,
in contrast to the nonpermanent or condensable gases like chlorine, nitrous oxide
and carbon dioxide, which could be liquefied at quite modest pressures of
30–50 atm. During the previous 50 years or so, in extremely dangerous experiments,

© 2005 by Marcel Dekker


5367-4 Flynn Ch01 R2 090704

Cryogenic Engineering Connections 7

a number of workers had discovered by visual observation in thick-walled glass tubes


that the permanent gases, including hydrogen, nitrogen, oxygen and carbon monox-
ide, could not be liquefied at pressures as high as 400 atm. The success of these
experimenters marked the end of the idea of permanent gases and established the
possibility of liquefying any gas by moderate compression at temperatures below
the critical temperature. In 1866, Van der Waals (1837–1923), had published his first
paper on ‘‘the continuity of liquid and gaseous states’’ from which the physical
understanding of the critical state, and of liquefaction and evaporation, was to grow.
Cailletet had used the apparatus shown in Fig. 1.4 to produce a momentary fog
of oxygen droplets in the thick-walled glass tube. The oxygen gas was compressed
using the crude Natterer compressor in which pressures up to 200 atmospheres were
generated by a hand-operated screw jack. The pressure was transmitted to the oxy-
gen gas in the glass tube by hydraulic transmission using water and mercury. The gas
was cooled to 103 C by enclosing the glass tube with liquid ethylene and was then
expanded suddenly by releasing the pressure via the hand wheel. A momentary fog
was seen, and the procedure could then be repeated for other observers to see the
phenomenon.
Figure 1.5 shows the cascade refrigeration system used by Professor Pictet at
the University of Geneva in which oxygen was first cooled by sulphur dioxide and
then by liquid carbon dioxide in heat exchangers, before being expanded into the
atmosphere by opening a valve. The isenthalpic expansion yielded a transitory jet of
partially liquefied oxygen, but no liquid could be collected from the high-velocity
jet. The figure shows how Pictet used pairs of compressors to drive the SO2 and
CO2 refrigerant cycles. This is probably one of the first examples of the cascade
refrigeration system invented by Tellier (1866) operating at more than one tempera-
ture level. Pictet was a physicist with a mechanical flair, and although he did not

Figure 1.4 Cailletet’s gas compressor and liquefaction apparatus (1877).

© 2005 by Marcel Dekker


5367-4 Flynn Ch01 R2 090704

8 Chapter 1

Figure 1.5 Pictet’s cascade refrigeration and liquefaction system (1877).

pursue the liquefaction of oxygen (he made a name developing ice-skating rinks), his
use of the cascade system inspired others like Kamerlingh Onnes and Dewar.
In the early 1880s one of the first low-temperature physics laboratories, the
Cracow University Laboratory in Poland, was established by Szygmunt von
Wroblewski and K. Olszewski. They obtained liquid oxygen ‘‘boiling quietly in a test
tube’’ in sufficient quantity to study properties in April 1883. A few days later, they
also liquefied nitrogen. Having succeeded in obtaining oxygen and nitrogen as true
liquids (not just a fog of liquid droplets), Wroblewski and Olszewski, now working
separately at Cracow, attempted to liquefy hydrogen by Cailletet’s expansion tech-
nique. By first cooling hydrogen in a capillary tube to liquid-oxygen temperatures
and expanding suddenly from 100 to 1 atm, Wroblewski obtained a fog of liquid-
hydrogen droplets in 1884, but he was not able to obtain hydrogen in the completely
liquid form.
The Polish scientists at the Cracow University Laboratory were primarily inter-
ested in determining the physical properties of liquefied gases. The ever-present prob-
lem of heat transfer from ambient plagued these early investigators because the
cryogenic fluids could be retained only for a short time before the liquids boiled
away. To improve this situation, an ingenious experimental technique was developed
at Cracow. The experimental test tube containing a cryogenic fluid was surrounded
by a series of concentric tubes, closed at one end. The cold vapor arising from the
liquid flowed through the annular spaces between the tubes and intercepted some
of the heat traveling toward the cold test tube. This concept of vapor shielding is used
today in conjunction with high-performance insulations for the long-term storage of
liquid helium in bulk quantities.
All over Europe scientists worked to produce a system that would operate to
make liquid gas on an industrial scale. The major problem in all this was to prevent

© 2005 by Marcel Dekker


5367-4 Flynn Ch01 R2 090704

Cryogenic Engineering Connections 9

the material from drawing heat from its surroundings. In 1882, a French physicist
called Jules Violle wrote to the French Academy to say that he had worked out a
way of isolating the liquid gas from its surroundings through the use of a vacuum.
It had been known for some time that vacua would not transmit heat, and Violle’s
arrangement was to use a double-walled glass vessel with a vacuum in the space
between the walls. Violle has been forgotten, his place taken by a Scotsman who
was to do the same thing, much more efficiently, eight years later. His name was
Sir James Dewar, and he added to the vessel by silvering it both inside and out
(Violle had only silvered the exterior), in order to prevent radiation of heat either
into or out of the vessel.

6. THE SCIENTISTS

James Dewar was appointed to the Jacksonian Professorship of Natural Philosophy


at Cambridge in 1875 and to the Fullerian Professorship of Chemistry at The Royal
Institution in 1877, the two appointments being held by him until his death at the age
of 81 in 1923. Dewar’s research interests ranged widely but his outstanding work was
in the field of low temperatures. Within a year of his taking office at The Royal Insti-
tution, the successful liquefaction of oxygen by Cailletet and Pictet led Dewar to
repeat Cailletet’s experiment. He obtained a Cailletet apparatus from Paris, and,
within a few months, in the summer of 1878 he demonstrated the formation of a mist
of liquid oxygen to an audience at one of his Friday evening discourses at The Royal
Institution. This lecture was the first of a long series of demonstrations, extending
over more than 30 years and culminating in dramatic and sometimes hazardous
demonstrations with liquid hydrogen (Fig. 1.6). In May 1898 Dewar produced
20 cm3 of liquid hydrogen boiling quietly in a vacuum-insulated tube, instead of
a mist.
The use of a vacuum had been used by Dewar and others as early as 1873 and
his experiments over several years before 1897 went on to show how he could obtain
significant reductions (up to six times) by introducing into the vacuum space pow-
ders such as charcoal, lamp black, silica, alumina, and bismuth oxide. For this
purpose, he used sets of three double-walled vessels connected to a common vacuum
in which one of the set was used as a control. Measurement of the evaporation rate
of liquid air in the three vessels then enabled him to make comparative assessments
on the test insulations.
In 1910 Smoluchowski demonstrated the significant improvement in insulating
quality that could be achieved by using evacuated powders in comparison with unevac-
uated insulations. In 1937, evacuated-powder insulations were first used in the United
States in bulk storage of cryogenic liquids. Two years later, the first vacuum-powder-
insulated railway tank car was built for the transport of liquid oxygen.
Following evacuated powders, he made further experiments using metallic and
other septa-papers coated with metal powders in imitation of gold and silver,
together with lead and aluminum foil and silvering of the inner surfaces of the annu-
lar space. He found that three turns of aluminum sheet (not touching) were not as
good as silvered surfaces. Had he gone on to apply further turns of aluminum, he
would have discovered the principle of multilayer insulation which we now know
to be superior to silvering. Nevertheless, his discovery of silvering as an effective
means of reducing the radiative heat flux component was a breakthrough. From
1898, the glass Dewar flask quickly became the standard container for cryogenic

© 2005 by Marcel Dekker


5367-4 Flynn Ch01 R2 090704

10 Chapter 1

Figure 1.6 Sir James Dewar lecturing at the Royal Institution. Although Violle preceded
Dewar in the development of the vacuum flask, there is no evidence that Dewar knew of
his work when he presented the details of his new container to the Royal Institution in 1890.

liquids. Meanwhile his work on the absorptive capacity of charcoal at low tempera-
tures paved the way towards the development of all-metal, double-walled vacuum
vessels.
Another first for Dewar was his use of mixtures of gases to enhance J-T
cooling, a topic revisited in cryogenics in 1995.
Historically, the expansion of a mixture of hydrogen and nitrogen was
employed by Dewar in 1894 in attempts to liquefy hydrogen at that time. Dewar
wrote: ‘‘Expansion into air at one atmosphere pressure of a mixture of 10% nitrogen
in hydrogen yielded a much lower temperature than anything that has been recorded
up to the present time.’’
Because his flasks (unlike those of Jules Violle) were silvered inside and out,
Dewar’s flasks could equally well retain heat as cold. Dewar’s vessel became known
in scientific circles as the Dewar flask; with it, he was able to use already liquid gases

© 2005 by Marcel Dekker


5367-4 Flynn Ch01 R2 090704

Cryogenic Engineering Connections 11

to enhance the chill during the liquefaction of gases whose liquefaction temperature
was lower than that of the surrounding liquid. In this way, in 1891, he succeeded for
the first time in liquefying hydrogen.
Dewar had considerable difficulty in finding competent glass makers willing to
undertake the construction of his double-walled vessels and had been forced to get
them made in Germany.
By 1902 a German called Reinhold Burger, whom Dewar had met when visit-
ing Germany to get his vessels made, was marketing them under the name of
Thermos. The manufacture of such vessels developed into an important industry
in Germany as the Thermos flask, and this monopoly was maintained up to 1914.
Dewar never patented his silvered vacuum flask and therefore never benefited finan-
cially from his invention.
The word ‘‘Cryogenics’’ was slow in coming.
The word cryogenics is a product of the 20th Century and comes from the
Greek—kroB—frost and—ginomai—to produce, engender. Etymologically, cryo-
genics means the science and art of producing cold and this was how Kamerlingh
Onnes first used the word in 1894. Looking through the papers and publications
of Dewar and Claude, it appears that neither of them ever used the word; indeed
a summary of Dewar’s achievements by Armstrong in 1916 contains no mention
but introduces yet another term, ‘‘the abasement of temperature.’’
In 1882, Kamerlingh Onnes (1853–1926) was appointed to the Chair of Experi-
mental Physics at the University of Leiden in the Netherlands and embarked on
building up a low-temperature physics laboratory in the Physics Department. The
inspiration for his laboratory was provided by the systematic work of Van der Waals
at Amsterdam, and subsequently at Leiden, on the properties of gases and liquids. In
1866, Van der Waals had published his first paper on ‘‘the continuity of liquid and
gaseous states’’ from which the physical understanding of the critical state, and of
liquefaction and evaporation was to grow.
He operated an open-door policy encouraging visitors from many countries to
visit, learn, and discuss. He published all the experimental results of his laboratory,
and full details of the experimental apparatus and techniques developed, by introdu-
cing in 1885 a new journal ‘‘Communications from the Physical Laboratory at the
University of Leiden.’’
As a result, he developed a wide range of contacts and a growing track record
of success, so that from the turn of the century his Leiden laboratory held a leading
position in cryogenics for almost 50 years, certainly until the mid-1930s. In 1908, for
example, he won the race with Dewar and others to liquefy helium and went on to
discover superconductivity in 1911.
Onnes’ first liquefaction of helium in 1908 was a tribute both to his experimen-
tal skill and to his careful planning. He had only 360 L of gaseous helium obtained
by heating monazite sand from India. More than 60 cm3 of liquid helium was pro-
duced by Onnes in his first attempt. Onnes was able to attain a temperature of
1.04 K in an unsuccessful attempt to solidify helium by lowering the pressure above
a container of liquid helium in 1910.
It is interesting to compare Dewar’s approach with that of his rival,
Kamerlingh Onnes. Having successfully liquefied hydrogen in 1898, Dewar had been
able to monopolize the study of the properties of liquid hydrogen and published
many papers on this subject. His attempts in 1901 to liquefy helium in a Cailletet
tube cooled with liquid hydrogen at 20.5 K, using a single isentropic expansion from
pressures up to 100 atm, led at first to a mist, being clearly visible. Dewar was

© 2005 by Marcel Dekker


5367-4 Flynn Ch01 R2 090704

12 Chapter 1

suspicious of a contamination because after several compressions and expansions,


the end of the Cailletet tube contained a small amount of solid that sublimed to a
gas without passing through the liquid state when the liquid hydrogen was removed.
On lowering the temperature of the liquid hydrogen by pumping to 16 K, and repeat-
ing the expansions on the gas from which the solid had been separated by the
previous expansions at 20.5 K, no mist was seen. From these observations,
he concluded that the mist was caused by some material other than helium, in all
probability neon, and that the critical temperature and boiling point of helium
were below 9 K and about 5 K, respectively.
The Cailletet tube was, of course, limited in its potential and Dewar appre-
ciated that he needed a continuous circulation cascade system employing liquid
hydrogen at reduced pressure and a final stage of Joule–Thomson expansion with
recuperative cooling. He already had at the Royal Institution a large quantity of
hardware, compressors, and pumps for the cascade liquefaction system he had
assembled for liquefying hydrogen. From 1901, he joined the race to liquefy helium
with competitors like Travers and Ramsay at University College, London,
Kamerlingh Onnes and his Leiden team, and Olszewski at Cracow. The race
continued for 7 years until 1908 when the Leiden team won. Dewar and the other
competitors perhaps failed because they had not appreciated that the magnitude
of the effort to win the race required a systems approach to solve the problems of
purification, handling small quantities of precious helium gas, maintaining leaky
compressors, improving the design of recuperative heat exchangers, and under-
standing the properties and behavior of fluids and solids at low temperatures, all
at the same time.
As a direct result of their success, the Leiden team of Kamerlingh Onnes went
on to discover superconductivity in 1911 and thereafter maintained a leading
position in low-temperature research for many years.

7. THE ENGINEERS

Dr. Hampson was medical officer in charge of the Electrical and X-ray Departments,
Queens and St. John’s Hospitals, Leicester Square, London. He was a product of the
Victorian age, with a classics degree at Oxford in 1878, having subsequently acquired
his science as an art and a living; but he possessed an extraordinary mechanical flair.
He was completely overshadowed by Dewar at The Royal Institution, Ramsay at
University College and Linde at Wiesbaden, each with their considerable laboratory
facilities. Indeed, Dewar seems to have been unable to accept Hampson as a fellow
experimentalist or to acknowledge his contribution to cryogenics.
And yet, Hampson with his limited facilities was able to invent and develop a
compact air liquefier which had a mechanical elegance and simplicity which made
Dewar’s efforts crude and clumsy in comparison. Indeed, Hampson’s design of heat
exchanger was so successful that it is still acknowledged today. The Hampson type
coiled tube heat exchanger is widely used today. In 1895, when Hampson and Linde
independently took out patents on their designs of air liquefier, the Joule–Thomson
effect was known and the principle of recuperative cooling by so-called self-
intensification had been put forward by Siemens as early as 1857. The step
forward in both patents was to break away from the cascade system of cooling
and to rely entirely on Joule–Thomson cooling together with efficient heat exchanger
designs.

© 2005 by Marcel Dekker


5367-4 Flynn Ch01 R2 090704

Cryogenic Engineering Connections 13

Figure 1.7 Linde two-stage compressor and 3 L=hr air liquefier (1895).

Linde used concentric tubes of high-pressure line enclosed by a low-pressure


return, the two being wound into a single layer spiral to achieve a compact design
(Fig. 1.7). ‘‘Compact’’ is a relative term. Linde’s first such heat exchanger was made
of hammered copper tubing approximately one-quarter of an inch thick. The
outer (low pressure) tube was more than 6 in. in diameter. It is said that it took
the better part of a month for the heat exchanger to cool down and achieve thermal
equilibrium.
After 1895, Linde made rapid progress in developing his Joule–Thomson
expansion liquefier making some 3 L of liquid air per hour.
By the end of 1897, Charles Tripler, an engineer in New York had constructed
a similar but much larger air liquefier, driven by a 75 kW steam engine, which pro-
duced up to 15 L of liquid air per hour (Fig. 1.8). Tripler discovered a market for
liquid air as a medium for driving air expansion engines—the internal combustion
engine was still unreliable at that time—and he succeeded in raising $10 M on Wall
Street to launch his Liquid Air Company.
Using the liquid air he had produced to provide high-pressure air for an air
expansion engine to drive his air compressor, he was convinced that he could make
more liquid air than he consumed—and coined the word ‘‘surplusage.’’ He was of
course wrong. In 1902, Tripler was declared bankrupt, and Wall Street and the
US lost interest in commercial applications of cryogenics for many years to come,
although important cryophysics and cryo-engineering research continued in US
universities.
In 1902 Georges Claude, a French engineer, developed a practical system for
air liquefaction in which a large portion of the cooling effect of the system was
obtained through the use of an expansion engine. The use of an expansion engine

© 2005 by Marcel Dekker


5367-4 Flynn Ch01 R2 090704

14 Chapter 1

Figure 1.8 Tripler’s air liquefier with steam-driven compressor and tube-in-shell heat
exchangers (1898).

caused the gas to cool isentropically rather than isenthalpically, as is the case in
Joule–Thomson expansion. Claude’s first engines were reciprocating engines using
leather seals (actually, the engines were simply modified steam engines). During
the same year, Claude established l’Air Liquide to develop and produce his systems.
The increase in cooling effect over the Joule–Thomson expansion of the
Linde=Hampson=Tripler designs was so large as to constitute another technical
breakthrough. Claude went on to develop air liquefiers with piston expanders in
the newly formed Société L’Air Liquide.
Although cryogenic engineering is considered a relatively new field in the US, it
must be remembered that the use of liquefied gases in US industry began in the early
1900s. Linde installed the first air-liquefaction plant in the United States in 1907, and
the first American-made air-liquefaction plant was completed in 1912. The first com-
mercial argon production was put into operation in 1916 by the Linde company in
Cleveland, Ohio. In 1917 three experimental plants were built by the Bureau of
Mines, with the cooperation of the Linde Company, Air Reduction Company,
and the Jefferies-Norton Corporation, to extract helium from natural gas of Clay
County, Texas. The helium was intended for use in airships for World War I.
Commercial production of neon began in the United States in 1922, although Claude
had produced neon in quantity in France since 1920.
Around 1947 Dr. Samuel C. Collins of the department of mechanical engineer-
ing at Massachusetts Institute of Technology developed an efficient liquid-helium
laboratory facility. This event marked the beginning of the period in which liquid-
helium temperatures became feasible and fairly economical. The Collins helium
cryostat, marketed by Arthur D. Little, Inc., was a complete system for the safe,
economical liquefaction of helium and could be used also to maintain temperatures
at any level between ambient temperature and approximately 2 K.

© 2005 by Marcel Dekker


5367-4 Flynn Ch01 R2 090704

Cryogenic Engineering Connections 15

The first buildings for the National Bureau of Standards Cryogenic Engineering
Laboratory were completed in 1952. This laboratory was established to provide engi-
neering data on materials of construction, to produce large quantities of liquid hydro-
gen for the Atomic Energy Commission, and to develop improved processes and
equipment for the fast-growing cryogenic field. Annual conferences in cryogenic engi-
neering have been sponsored by the National Bureau of Standards (sometimes spon-
sored jointly with various universities) from 1954 (with the exception of 1955) to 1973.
At the 1972 conference at Georgia Tech in Atlanta, the Conference Board voted to
change to a biennial schedule alternating with the Applied Superconductivity Confer-
ence. The NBS Cryogenic Engineering Laboratory is now part of history, as is the
name ‘‘NBS’’, now the National Institutes for Science and Technology (NIST).

8. THE ROCKET SCIENTISTS

The impact of cryogenics was wide and varied. The Dewar flask changed the social
habits of the Edwardian well-to-do: picnics became fashionable because of it. In time
it changed the working-man’s lunch break and accompanied expeditions to the tropics
and to the poles, carrying sustenance for the explorers and returning with hot or cold
specimens. Later, it saved thousands of lives by keeping insulin and other drugs from
going bad. Perhaps its most spectacular impact was made, however, by two men whose
work went largely ignored, and by a third who did his work in a way that could not be
ignored. The first was a Russian called Konstantin Tsiolkovsky, whose early use of
liquid gas at the beginning of the 20th century was to lie buried under governmental
lack of interest for decades. The second was an American called Robert Goddard,
who did most of his experiments on his aunt’s farm in Massachusetts, and whose only
reward was lukewarm interest from the weather bureau.
On 16 March 1926, Dr. Robert H. Goddard conducted the world’s first suc-
cessful flight of a rocket powered by liquid-oxygen–gasoline propellant on a farm
near Auburn, Massachusetts. This first flight lasted only 212 sec, and the rocket
reached a maximum speed of only 22 m=s (50 mph). Dr. Goddard continued his
work during the 1930s, and by 1941 he had brought his cryogenic rockets to a fairly
high degree of perfection. In fact, many of the devices used in Dr. Goddard’s rocket
systems were used later in German V-2 weapons systems (Fig 1.9).
The third was a German, Herman Oberth, and his work was noticed because it
aimed at the destruction of London.
His liquid gases were contained in a machine that became known as Vengeance
Weapon 2, or V-2, and by the end of the Second World War it had killed or injured
thousands of military and civilian personnel. All three men had realized that certain
gases burn explosively, in particular hydrogen and oxygen, and that, since in their
liquid form they occupy less space than as a gas (hydrogen does this by a factor
of 790) they were an ideal fuel. Thanks to the principles of the Dewar flask, they
could be stored indefinitely, transported without loss, and contained in a launch
vehicle that was essentially a vacuum flask with pumps, navigation systems, a
combustion chamber and a warhead.
(One of Oberth’s most brilliant assistants was a young man called Werner von
Braun, and it was he who brought the use of liquid fuel to its most spectacular
expression when his brainchild, the Saturn V, lifted off at Cape Canaveral on 16 July
1969, carrying Armstrong, Aldrin, and Collins to their historic landing on the
moon.)

© 2005 by Marcel Dekker


5367-4 Flynn Ch01 R2 090704

16 Chapter 1

Figure 1.9 The V-2 liquid-fueled rocket used a mixture of oxygen and kerosene. Originally
developed at the experimental rocket base in Peenemunde, on the Baltic, the first V-2 landed in
London in 1944. When the war ended, German engineers were working on a V-3 capable of
reaching New York.

There is no doubt that the development of today’s space technology would


have been impossible without cryogenics. The basic reason for this lies with the high
specific impulse attainable with kerosene=liquid oxygen (2950 m=s) and liquid hydro-
gen=liquid oxygen (3840 m=s)—values much higher than with liquid or solid propel-
lants stored at ambient temperatures.
Space cryogenics developed rapidly in the early 1960s for the Apollo rocket ser-
ies, at the same time as LNG technology. The driving force for space cryogenics
development was the competition between the US and Russia in the exploration
of Space, and the surface of the Moon in particular, and in the maintenance of
detente in the ‘‘cold war’’.
Particular requirements then, and now in the Space Shuttle flight series, include
the use of liquid hydrogen as a propellant fuel, liquid oxygen as a propellant oxidizer
and for the life support systems, both liquids for the fuel cell electric power supplies
and liquid helium to pressurize the propellant tanks. The successful development of
the necessary cryogenic technology has provided an extraordinary range of spin-offs
and a remarkable level of confidence in the design, construction, and handling of
cryogenic systems.

9. THE PHYSICISTS AND SUPERCONDUCTIVITY

The phenomenon of zero electrical resistance was discovered in 1911 in mercury


by Kamerlingh Onnes and his team at Leiden. Although he realized the great

© 2005 by Marcel Dekker


5367-4 Flynn Ch01 R2 090704

Cryogenic Engineering Connections 17

significance of his discovery, Kamerlingh Onnes was totally frustrated by the lack of
any practical application because he found that quite small magnetic fields, applied
either externally or as self-fields from internal electric currents, destroyed the
superconducting state. The material suddenly quenched and acquired a finite electri-
cal resistance.
For the next 50 years, very little progress was made in applying superconduc-
tivity, although there was a growing realization that a mixed state of alternate lami-
nae of superconducting and normal phase had a higher critical field which depended
on the metallurgical history of the sample being studied. In fact, an impasse had
arisen by the late 1950s, and little progress was being achieved on the application
of superconductivity.
On the other hand, great progress had been achieved by this time in the UK,
USSR, and US in the theoretical description of superconductivity. This progress
stemmed largely from two sets of experimental evidence; firstly, demonstrations of
the isotope effect, indicating that lattice vibrations must play a central role in the
interaction leading to superconductivity; secondly, the accumulation of evidence that
an energy gap exists in the spectrum of energy states available to the conduction elec-
trons in a superconductor. In 1956, Bardeen, Cooper, and Schrieffer (see Table 1.1.
for a list of the many Nobel Laureates indebted to cryogenics) proposed a successful
theory of superconductivity in which conduction electrons are correlated in pairs
with the same center of mass momentum via interactions with the lattice. The theory
made predictions in remarkably good agreement with experimental data and
provided the basis for later developments such as quantum mechanical tunneling,
magnetic flux quantization, and the concept of coherence length.
The break out of the impasse in the application of superconductivity came from
systematic work at the Bell Telephone Laboratories, led by Matthias and Hulm. In
1961, they published their findings on the brittle compound Nb3Sn, made by the
high-temperature treatment of tin powder contained in a Niobium capillary tube. This
compound retained its superconductivity in a field of 80 KG (8 Tesla), the highest field
available to them, while carrying a current equivalent to a density of 100,000 A=cm2.
The critical field at 4.2 K, the boiling point of helium, was much greater than 8 T and
was therefore at least 100 times higher than that of any previous superconductor.
This finding was followed by the discovery of a range of compounds and alloys
with high critical fields, including NbZr, NbTi, and V3Ga; the race was on to man-
ufacture long lengths of wire or tape and develop superconducting magnets for
commercial applications. At first, the early wires were unstable and unreliable and
it soon became clear that a major research effort was needed.

9.1. High Field, Type 2 Superconductivity (1961)


In any event, NbTi turned out to be a much easier material to develop than Nb3Sn,
because it was ductile. However, it took 10 years or more to develop reliable, intern-
ally stabilized, multifilamentary composite wires of NbTi and copper. Only then,
around 1970, was it possible for high field and large-scale applications of super-
conductivity to be considered with confidence.

9.2. The Ceramic Superconductors


All previous developments in superconductivity were eclipsed in 1986 by the discov-
ery of a new type of superconductor composed of mixtures of ceramic oxides. By the

© 2005 by Marcel Dekker


5367-4 Flynn Ch01 R2 090704

18 Chapter 1

Table 1.1 Nobel Laureates Linked with Cryogenics

1902 Pieter Zeeman Influence of magnetism upon radiation


1904 Lord Rayleigh Density of gases and his discovery of argon
1910 Johannes Van der Waals The equation of state of gases and liquids
1913 H. Kamerlingh Onnes The properties of materials at low
temperature, the preparation of liquid
helium
1920 Charles Guillaume Materials for national prototype standards
(Ni-Steel) INVAR
1934 H. C. Urey Discovery of deuterium produced by the
distillation of liquid hydrogen
1936 P.J.W. Debye The behavior of substances at extremely low
temperatures, especially heat capacity
1950 Emmanuel Maxwell ‘‘Isotope Effect’’ in superconductivity
1956 John Bardeen Semiconductors and superconductivity
1957 Tsung Dao Lee, Upsetting the principle of conservation of
Chen Ning Yang parity as a fundamental law of physics.
1960 D. A. Glaser Invention of bubble chamber
1961 R. L. Mössbauer Recoil-less nuclear resonance absorption of
gamma radiation
1962 L. D. Landau Pioneering theories of condensed matter
especially liquid He3
1968 Louis W. Alvarez Decisive contributions to elementary
particle physics, through his development
of hydrogen bubble chamber technique
and data analysis
1972 John Bardeen, Leon N. Cooper, BCS theory of superconductivity
J. Robert Schrieffer
1973 B. D. Josephson, Ivar Giaever, The discovery of tunneling supercurrents
Leo Esaki (The Josephson Junction)
1978 Peter Kapitza Methods pf making liquid helium and the
characterization of Helium II as a
‘‘superfluid’’.
1987 J. G. Bednorz, K.A. Muller High-temperature superconductors
1996 D. Lee, D. D. Osheroff, Discovery of superfluidity in liquid helium-3
R.C. Richardson
1997 Steven Chu, Methods to cool and trap atoms with lasers
Claude Cohen-Tannoudji
1998 R. B. Laughlin, H. L. Stormer, Discovery of a new form of quantum fluid
D. C. Tsui excitations at low temperatures
2001 Eric Cornell, W. Ketterle, For achieving the Bose–Einstein condensate
Carl Wieman at near absolute zero

end of 1987, superconducting critical temperatures of ceramic materials had risen to


90 K for Y–Ba–Cu–O; 110 K for Bi–Sr–Ga–Cu–O; and 123 K for Tl–Ba–Ca–Cu–O.
These temperatures are about ten times higher than the critical temperatures of pre-
vious metallic superconductors, thereby allowing liquid nitrogen instead of liquid
helium to be used as a cooling medium.
This discovery changes the economics and practicability of engineering appli-
cations of superconductivity in a dramatic way. Refrigeration costs with liquid nitro-
gen are 100 times cheaper than those with liquid helium, and only simple one-stage

© 2005 by Marcel Dekker


5367-4 Flynn Ch01 R2 090704

Cryogenic Engineering Connections 19

Table 1.2 Notable Events in the History of Cryogenics

1848 John Gorrie produces first mechanical refrigeration machine


1857 Siemens suggests recuperative cooling or ‘‘self-intensification’’
1866 Van der Waals first explored the critical point, essential to the work of Dewar
and Onnes
1867 Henry P. Howard of San Antonio Texas uses Gorrie’s air-chilling system to
transport frozen beef from Indianola TX to cities along the Gulf of Mexico
1869 Malthusian revolution in England, predicting worldwide starvation
1873 James Harrison attempts to ship frozen beef from Australia to the UK aboard
the SS Norfolk., the project failed
1875 Thomas Mort tries again to ship frozen meat from Australia to the UK, this time
aboard the SS Northam. Another failure. Gorrie’s air system eventually
produce success by Bell and Coleman 1877
1877 Coleman and Bell produce commercial version of Gorrie’s system for freezing
beef. The frozen meat trade becomes more successful and stems the Malthusian
revolution
Cailletet produced a fog of liquid air, and Pictet a jet of liquid oxygen
1878 James Dewar duplicates the Cailletet=Pictet experiment before the Royal
Institution
1879 The SS. Strathaven arrives in London carrying a well preserved cargo of frozen
meat from Australia
Linde founded the Linde Eismachinen
1897 Charles Tripler of NY produces 15 L=hr of liquid air using a 75 kW steam engine
power source. Liquid air provides high-pressure gas to drive his air compressor
and tripler is convinced he can make more liquid air than he consumes, the
‘‘surplusage’’ effect
1882 Jules Violle develops the first vacuum insulated flask
1883 Wroblewski and Olszewski liquefied both nitrogen and oxygen
Vapor cooled shielding developed by Wroblewski and Olszewski
1884 Wroblewski produced a mist of liquid hydrogen
1891 Linde had put 12,000 domestic refrigerators in service.
Dewar succeeds for the first time in liquefying hydrogen in a mist
1892 Dewar developed the silvered, vacuum-insulated flask that bears his name
1894 Dewar first demonstrated benefit of gas mixtures in J-T expansion
Onnes first uses the word ‘‘cryogenics’’ in a publication
1895 Kammerlingh Onnes established the University of Leiden cryogenics laboratory
Linde is granted the basic patent on air liquefaction
First Hampson heat exchanger made for air liquefaction plants
Hampson and Linde independently patent air liquefiers using Joule–Thomson
expansion and recuperative cooling
1897 Dewar demonstrates the vacuum powder insulation
1898 Dewar produced liquid hydrogen in bulk, at the royal Institute of London
1890 Dewar improves upon the Violle vacuum flask by slivering both surfaces
1902 Claude developed an air liquefaction system using an expansion engine using
leather seals, and established l’Air Liquide
Reinhold Burger markets Dewar vessels under the name Thermos2
Tripler (1897) files for bankruptcy after his ‘‘surplusage’’ is proven false
1907 Linde installed first air liquefaction plant in the USA
Claude produced neon as a by product of an air plant
1908 Onnes liquefied helium and received the Noble prize for his accomplishment
1910 Smoluchowski demonstrated evacuated-powder insulation
Van der Waals receives Nobel prize for work on the critical region

(Continued)

© 2005 by Marcel Dekker


5367-4 Flynn Ch01 R3 101104

20 Chapter 1

Table 1.2 (Continued )

1911 Onnes discovered superconductivity


1912 First American made air liquefaction plant
1913 Kammerlingh Onnes receives Nobel for work on liquid Helium
1916 First commercial production of argon in the USA by the Linde Company
1917 First natural gas plant produces gaseous helium
1920 Commercial production of neon in France
1922 First commercial neon production in the USA
1926 Dr. Goddard fired the first cryogenically propelled rocket
Giauque and Debye independently discover the adiabatic demagnetization
principle for producing temperatures much lower than 1 K
1933 Magnetic cooling first produces temperatures below 1 K
1934 Kapitza produces first expansion engine for making liquid helium
H. C. Urey receives the Nobel for his discovery of deuterium
1936 P.J.W. debye receives Nobel prize for discoveries on heat capacity at low
temperatures
1937 Evacuated-powder insulation, originally tested by Dewar, is first used on a
commercial scale in cryogenic storage vessels
First vacuum-insulated railway tank car built for liquid oxygen
Hindenburg Zeppelin crashed and burned at Lakehurst, NJ. The hydrogen
burned but did not explode. Thirty-seven out of 39 passengers survived, making
this disaster eminently survivable as far as air travel goes (63 out of 100 total
passengers and crew survived). Nonetheless, hydrogen gains an extremely bad
reputation
1942 V-2 liquid oxygen fueled weapon system fired
1944 LNG tank in Cleveland OH fails killing 131 persons. LNG industry set back 25
years in the USA
1947 Collins cryostat developed making liquid helium readily available for the first
time
1948 First 140 ton=day oxygen system built in the USA
1950 Emmanuel Maxwell receives Nobel prize for discovering the isotope effect in
superconductors
1952 National Bureau of Standards Cryogenic Engineering laboratory built in
Boulder, Colorado, including the first large-scale liquid hydrogen plant in the
USA
1954 First Cryogenic Engineering Conference held by NBS at its Boulder Laboratories
1956 BCS theory of superconductivity proposed
1957 Atlas ICBM tested, firs use in the USA of liquid oxygen-RP1 propellant
Lee and Yang receive Nobel for upsetting the theory of parity
1958 Multilayer cryogenic insulation developed
1959 The USA space agency builds a tonnage liquid hydrogen plant at Torrance, CA
1960 D.A. Glaser receives Nobel for his invention of the bubble chamber
1961 Liquid hydrogen fueled Saturn launch vehicle test fired
R. L. Mossbauer receives Nobel for discoveries in radiation absorption
1962 L. D. Landau awarded the Nobel Prize for discoveries in He3
1968 L. W. Alvarez receives Nobel Prize for his work with liquid hydrogen bubble
chambers
1969 Liquid hydrogen fueled Saturn vehicle launched from Cape Canaveral, FL
carrying Armstrong, Aldrin, and Collins to the moon
1972 Bardeen, Cooper, and Schrieffer receive Nobel for the BCS theory of
superconductivity

(Continued)

© 2005 by Marcel Dekker


5367-4 Flynn Ch01 R2 090804

Cryogenic Engineering Connections 21

Table 1.2 (Continued )

1973 B. D. Josephson, I. Giaever, and L. Esaki awarded the Nobel prize for discovery
of the Josephson Junction (tunneling supercurrents)
1978 Peter Kapitza receive the Nobel for the characterization of HeII as a superfluid
1987 J. G. Bednorz and K. A. Mueller awarded the Nobel Prize for discovering
high-temperature superconductors
Y–BA–Cu–O ceramic superconductors found
1996 D. Lee, D. D. Osheroff, and R. C. Richardson receive the Nobel prize of the
discovery of superfluidity in helium-3
1997 S. Chu and Claude Cohen-Tannoudji awarded the Nobel for discovering methods
to cool and trap atoms with lasers
1998 R. B. Laughlin, H. L. Stormer, and D. C. Tsui receive the Nobel for discovering a
new form of quantum fluid excitations at extremely low temperatures
2001 E. Cornell, W. Ketterle, and C. Wieman awarded the Nobel for achieving the
Bose–Einstein condensate at near absolute zero

refrigerators are needed to maintain the necessary temperatures. Furthermore, the


degree of insulation required is much more simple, and the vacuum requirements
of liquid helium disappear.

10. SCIENCE MARCHES ON

Many Nobel laureates received their honors either because of work directly in cryo-
genics, or because of work in which cryogenic fluids were indispensable refrigerants.
Table 1.1 provides an imposing list of such Nobel Prize winners.
Two examples from this list of Nobel laureates are considered to illustrate the
application of cryogenics to education and basic research.
The first example involves Donald Glaser who invented the bubble cham-
ber. Glaser recognized the principal limitation of the Wilson cloud chamber,
namely that the low-density particles in the cloud could not intercept enough
high-energy particles that were speeding through it from the beams of powerful
accelerators. Glaser’s first bubble chamber operated near room temperature and
used liquid diethylether. To avoid the technical difficulties presented by such a
complex target as diethylether, Prof. Luis Alvarez, of the University of Califor-
nia, devised a bubble chamber charged with liquid hydrogen. Since hydrogen is
the simplest atom, consisting only of a proton and an electron, it interferes only
slightly with the high-energy processes being studied with the giant accelerators,
although it is readily ionized and serves quite well as the detector in the bubble
chamber.
The first really large liquid-hydrogen bubble chamber was installed at the
Lawrence Radiation Laboratory of the University of California and became opera-
tive in March 1959. A remark was made that this 72-in. liquid-hydrogen bubble
chamber would be the equivalent of a Wilson cloud chamber one-half mile long.
They could not really be equal, however, because the cloud chamber does not have
the great advantage of utilizing the simplest molecules as the detector.
The second example chosen from the list of Nobel laureates that serves to
link cryogenics to basic science is John Bardeen, who was cited for his work in

© 2005 by Marcel Dekker


5367-4 Flynn Ch01 R2 090704

22 Chapter 1

semiconductors, and more recently has worked in superconductivity. His exposition


in 1957, of what is now known as the BCS theory of superconductivity, has helped
make this field one of the most active research fields in physics in the last century.
Table 1.2 gives a chronology of some of the major events in the history of
cryogenics.

© 2005 by Marcel Dekker


5367-4 Flynn Ch02 R2 090804

2
Basic Principles

1. INTRODUCTION

In the last half of the 19th century, roughly from 1850 to the turn of the century, the
principal export of the United States was cotton. The second most important export
by dollar volume was ice. Americans developed methods for harvesting, storing, and
transporting natural ice on a global scale. In 1810, a Maryland farmer, Thomas
Moore, developed an ice-box to carry butter to market and to keep it hard until sold.
This early refrigerator was an oval cedar tub with an inner sheet of metal serving as a
butter container that could be surrounded on four sides by ice. A rabbit skin pro-
vided the insulated cover. Moore also developed an insulated box for home use. It
featured an ice container attached to the lid and a 6 cu ft storage space below. Ice
harvesting was revolutionized in 1825 when Nathaniel J. Wyeth invented a horse-
drawn ice cutter (see Fig. 2.1.) It was Wyeth’s method of cutting blocks of similar
size quickly and cheaply that made the bountiful supplies of natural ice resources
of the United States available for food preservation. A steam-powered endless chain
was developed in 1855 that could haul 600 tons of ice per hour to be stored. Wyeth
and Tudor patented a means to prevent the ice blocks from freezing together by pla-
cing saw-dust between the layers. Uniform blocks reduced waste, facilitated trans-
portation, and introduced ice to the consumer level (see Figs. 2.2 and 2.3).
The amount of ice harvested each year was staggering. In the winter of 1879–
1880, about 1,300,000 tons of ice was harvested in Maine alone, and in the winter of
1889–1890, the Maine harvest was 3 million tons. The Hudson River supplied about
2 million tons of ice per year to New York City during that period. Even that was
not enough, for during those same years, New York City imported about 15,000 tons
from Canada and 18,000 tons from Norway annually.
Ice changed the American diet. Fresh food preserved with ice was now pre-
ferred to food preserved by smoking or salting. Fresh milk could be widely distrib-
uted in the cities. By now, the people were accustomed to having ice on demand, and
great disruptions of the marketplace occurred when ice was not available. Ice was,
after all, a natural product, and a ‘‘bad’’ winter (a warm one) was disastrous to
the marketplace and to the diet. It was time for the invention of machines to make
ice on demand and free the market from dependence on the weather.
Machines employing air as a refrigerant, called compressed air or cold air
machines, were appearing. They were based on the reverse of the phenomenon of
heating that occurred when air was compressed, namely that air cooled as it
expanded against resistance. This phenomenon of nature had been observed as early
as the middle of the 18th century. Richard Trevithick, who lived until 1833 in
23

© 2005 by Marcel Dekker


5367-4 Flynn Ch02 R2 090804

24 Chapter 2

Figure 2.1 Ice harvesting and storage in the 1850s. Accumulated snow was first removed
with a horse-drawn ice plane. Then a single straight groove was made with a hand tool.
The ice was marked off from this line into squares 22 in. on a side. Actual cutting was begun
with Wyeth’s invention and finished with hand saws. The uniform blocks of ice were pulled
across the now open water to the lift at the icehouse. Here, they were stored under an insula-
tion of straw or sawdust. The availability of a relatively dependable supply of ice all year,
combined with uniformity of size of blocks, did much to create the market for ice.

Figure 2.2 Delivering ice in New York, 1884. In this neighborhood, apparently, icemen did
not enter the kitchen but sold their wares at the curb.

© 2005 by Marcel Dekker


5367-4 Flynn Ch02 R2 090804

Basic Principles 25

Figure 2.3 Commercial ice delivery in New York, 1884. We see delivery to a commercial
establishment, possibly a saloon. Though ice is available, the butcher a few doors down the
street still allows his meat to hang unrefrigerated in the open air.

Cornwall, England, constructed engines in which expanding air was used to convert
water to ice. In 1846, an American, John Dutton, obtained a patent for making ice
by the expansion of air. The real development of the cold air machine, however,
began with the one developed by John Gorrie of Florida and patented in England
in 1850 and in the United States in 1851.
Later, we shall see how Gorrie’s machine evolved into true cryogenics, not
merely ice production. For now, however, in the late 1800s, it was possible to buy
an ice-making machine for home use like that shown in Fig. 2.4. This primitive
refrigerator had the same components as any refrigerator of today. It also followed
exactly the same thermodynamic process as all of today’s refrigerators.
All refrigerators involve exchanging energy (work, as the young lady in Fig. 2.4
is doing) to compress a working fluid. The working fluid is later expanded against a
resistance, and the fluid cools (usually). Engineers have made it their job to deter-
mine how much cooling is produced, how much work is required for that cooling,
and what kind of compressor or other equipment is required. More than that, we
have made it our business to optimize such a cycle by trading energy for capital until
a minimum cost is found. We can make any process more ‘‘efficient’’, that is, less
energy-demanding, by increasing the size, complexity, and cost of the capital equip-
ment employed. For instance, an isothermal compressor requires less energy to
operate than any other kind of compressor. An isothermal compressor is also the
most expensive compressor to build, because an infinity of intercooling stages are
required for true isothermal compression.
As another example of trading energy costs for capital costs, consider the
Carnot cycle (described later), which is the most thermodynamically efficient heat
engine cycle possible. An essential feature of the Carnot engine is that all heat

© 2005 by Marcel Dekker


5367-4 Flynn Ch02 R2 090804

26 Chapter 2

Figure 2.4 Home ice machine. Machines such as this were available from catalog centers in
the late 1880s. The figure illustrates that energy (work) and capital equipment are at the heart
of all cryogenic processes.

transfer takes place with zero DT at the heat exchangers. As a consequence, a


Carnot engine must have heat exchangers of infinite size, which can get pretty costly.
In addition, the Carnot engine will never ‘‘get there’’ because another consequence
of zero DT is that an infinite amount of time will be required for heat transfer to
take place.

Someone noted for all posterity:


There once was a young man named Carnot
Whose logic was able to show
For a work source proficient
There is none so efficient
As an engine that simply won’t go.

The science that deals with these trade-offs between energy and capital is ther-
modynamics. Thermodynamics is about money. We need to predict how much a
refrigerator will cost to run (the energy bill) and how much it will cost to buy (the
capital equipment bill). Accordingly, our first job is to define heat and work, the
energy that we must buy. Work is fairly easily defined as the product of a force
andR a displacement. Its symbolic representation takes many different forms, such
as P dV, but it is always the product of an external force and the displacement that
goes with that force. Heat is more troublesome to calculate. The driving force for
heat, DT, seems obvious and analogous to the force or P term of the work equation.
But what is the equivalent displacement for heat? What gets moved, if there is such
an analogy?

© 2005 by Marcel Dekker


5367-4 Flynn Ch02 R2 090804

Basic Principles 27

Carnot reasoned, by the flow of water, that there must be a flow of heat. He
also reasoned that in the case of heat, something must be conserved, as energy itself
is conserved (here he was right). Carnot discovered that the quantity that was con-
served, and that was displaced, was not heat but rather the quantity Q=T. This led
Carnot to the concept of entropy. In this book, entropy (S) is used to calculate heat
(Q). That calculation is analogous to the one for work. Just as dW ¼ P dV, likewise
dQ ¼ T dS. There are certainly more sophisticated implications of entropy, but for
us, entropy is just a way to calculate how much heat energy is involved.
We need these ways to calculate heat and work because it is heat and work
(energy) that we pay for to run the equipment. You will never see a work meter
or a heat meter on any piece of cryogenic equipment. Instead it is a fact of nature
that these two economically important quantities must be calculated from the ther-
modynamically important quantities, the things that we can measure: pressure (P),
temperature (T), and density (actually, specific volume). Hence, we need two more
tools.
One essential tool is an equation of state to tie pressure, volume, and tempera-
ture together.
The second essential tool to help us figure out the cost is the thermodynamic
network that links not only P, V, and T together but also links all of the other ther-
modynamic properties (H, U, S, G, A, . . . —you know what I mean) together in a
concise and consistent way. In real life, we usually measure pressure and temperature
because it is possible to buy pressure gauges and thermometers and it is not possible
to buy entropy or enthalpy gauges. Hence, we need a way to get from P and T to H
and S. (The beautiful thermodynamic network that gets us around from P to T to
heat and work is shown later in Fig. 2.7.) This thermodynamic network is the only
reason we delve into partial differential equations, the Maxwell relations, and the
like. The payoff is that we can go from easily measured quantities, pressure and tem-
perature, to the hard-to-measure, but ultimately desired, quantities of heat and work.
You might want to look at Fig. 2.7 now to discover a reason to plow through all the
arcane stuff between here and there. Once we do get our hands on the thermody-
namic network, we can calculate heat and work for any process, for any fluid, for
all time.
We will start our cryogenic analysis with some definitions. Admittedly, this is a
boring approach. However, it will ensure that we are all singing from the same sheet
of music. We begin with nothing less than the definition of thermodynamics itself.

2. THERMODYNAMICS

Thermodynamics is built upon four great postulates (guesses), which are called,
rather stuffily, the four laws of thermodynamics.
The second law was discovered first; the first law was discovered second; the
third law is called the zeroth law; and the fourth law is called the third law. I am
sorry to tell you that things do not get much better.
The zeroth law of thermodynamics just says that the idea of temperature makes
sense.
The first law of thermodynamics is simply the conservation of energy principle.
Inelegantly stated, the first law says that what goes into a system must either come
out or accumulate. Undergraduates often call this the ‘‘checkbook law’’. If you can
balance your checkbook, you can do thermodynamics.

© 2005 by Marcel Dekker


5367-4 Flynn Ch02 R2 090804

28 Chapter 2

The second law is the entropy principle. The second law is also a conservation
equation, stating that in the ideal engine, entropy (S) is conserved. All real engines
make entropy.
The third law just says that there is a temperature so low that it can never be
reached.
These colloquial expressions of the four great principles of thermodynamics
have a calming effect. We will not be able to quantify our analyses unless we use
much more precise definitions and, expressly, the mathematical statements of these
principles. That is the task we must now turn to.
Thermodynamics is concerned with energy and its transformations into various
forms. The laws of thermodynamics are our concepts of the restrictions that nature
imposes on such transformations. These laws are primitive; they cannot be derived
from anything more basic. Unfortunately, the expressions of these laws use words
that are also primitive; that is, these words are in general use (e.g., energy) and have
no precise definition until we assign one. Accordingly, we begin our discussion of the
fundamental concepts of thermodynamics with a few definitions.
Thermodynamics: Thermodynamics is concerned with the interaction between a
system and its surroundings, the effect of that interaction on the properties of the
system, and the flow of heat and work between the system and its surroundings.
System: Any portion of the material universe set apart by arbitrarily chosen but
specific boundaries. It is essential that the system be clearly defined in any thermo-
dynamic analysis.
Surroundings: All parts of the material universe not included in the system.
The definitions of system and surroundings are coupled. The system is any
quantity of matter or region mentally set apart from the rest of the universe, which
then becomes the surroundings. The imaginary envelope that distinguishes the
system from the surroundings is called the boundary of the system.
Property of the system, or state variable: Any observable characteristic of the system,
such as temperature, pressure, specific volume, entropy, or any other distinguishing
characteristic.
State of the system: Any specific combination of all the properties of the system such
as temperature, pressure, and specific volume. Fixing any two of these three proper-
ties automatically fixes all the other properties of a homogeneous pure substance
and, therefore, determines the condition or state of that substance. In this respect,
thermodynamics is a lot like choosing the proper size dress shirt for a gentleman.
If you specify the neck size and sleeve length, the shirt size is defined. The size of
a man’s dress shirt is determined by two variables: neck size and sleeve length. A
single-phase homogeneous thermodynamic system is defined by fixing any two
thermodynamic variables. If you can pick out shirt size correctly, you can do
thermodynamics.
Energy: A very general term used to represent heat, work, or the capacity of a system
to do heat or work.
Heat: Energy in transition between a system and its surroundings that is caused to
flow by a temperature difference.
Work: Energy in transition between a system and its surroundings. Work is always
measured as the product of a force external to the system and a displacement of the
system.
If there is no system displacement, there is no work. If there is no energy in
transition, there is no work. Consider the following conundrum: It has been raining
steadily for the last hour at the rate of 1 in. of rain per hour. How much rain is on the

© 2005 by Marcel Dekker


5367-4 Flynn Ch02 R2 090804

Basic Principles 29

ground? Answer: none. Rain is water in transition between heaven and earth. Once
the water hits the earth, it is no longer rain. It may be a puddle or a lake, but it is not
rain.
Likewise, heat and work are energy in transition between system and surround-
ings. Heat and work exist only while in transition. Once transferred, heat and work
become internal energy, or increased velocity, or increased kinetic energy, but they
are no longer heat and work. Heat and work cannot reside in a system and are
not properties of a system.
Because heat and work are in transition, they have direction as well as magni-
tude, and there is a sign convention to tell us which way they are moving.
The convention among chemists and chemical engineers is:

Direction of energy flow Sign convention

Heat into the system (þ) Positive


Heat out of the system () Negative
Work into the system () Negative
Work out of the system (þ) Positive

Thus, the expression Q  W is the algebraic sum of all the energy flowing from
the surroundings into the system. Q  W is the net gain of energy of the system.
Process: The method or path by which the properties of a system change from one
set of values in an initial state to another set of values in a final state. For example, a
process may take place at constant temperature, at constant volume, or by any other
specified method. A process for which Q ¼ 0 is called an adiabatic process, for
instance.
Cyclic process: A process in which the initial state and final state are identical. The
Carnot cycle, mentioned earlier, is such a cyclic process. We will now examine the
Carnot cycle in the light of the preceeding definitions.
In 1800, there was an extraordinary Frenchman who was a statesman and gov-
ernment minister. His work was so important that he was known as the ‘‘organizer
of victory of the French Revolution’’. Lazare Nicholas Marguerite Carnot (1753–
1823) was also an outstanding engineer and scientist. A century later, this glorious
family tradition was carried on by his grandson, Marie François Sadi Carnot
(1837–1894), another gentleman engineer, who served as the President of the
Republic of France from 1887 to 1894, when he was assassinated.
The Carnot family member we are interested in is neither of these, but a cousin
who lived between them. Nicholas Leonard Sadi Carnot (1796–1832), called ‘‘Sadi’’
by everyone, was very interested in steam engines. He wanted to build the most effi-
cient steam engine of all to assist in military campaigns. Sadi needed the steam
engine that required the least fuel for a given amount of work, thereby reducing
the logistical nightmare of fueling the war engines.
He began by considering the waterwheel. He reasoned that it should be possi-
ble to hook two waterwheels together, one driven by falling water in the conven-
tional way. The other, joined to the first by a common shaft, should be able to
pick up the water spent by the first and carry it up to the headgate of the first. In
this perfect, frictionless engine, one waterwheel should be able to drive the other for-
ever. We know, as Sadi knew, that this is not possible because of the losses (friction)

© 2005 by Marcel Dekker


5367-4 Flynn Ch02 R2 090804

30 Chapter 2

in the system. However, the idea led Sadi to the conclusion that the most efficient
engine would be a reversible one, an engine that could run equally well in either
direction. The perfect waterwheel could be driven by falling water in one direction
or simply run in reverse to carry (pump) the water to the top again. A pair of such
reversible engines could thus run forever. Anything less than a reversible engine
would eventually grind to a halt. Hardly anyone paid any attention to Carnot and
his seminal concept, and indeed, Carnot’s work was further obscured by his early
death.
Carnot’s perfect abstraction was noticed, however, by the German physicist
Rudolph Clausius and the Glaswegian professor of natural philosophy William
Thomson. The principle of the conservation of energy was well established by this
time. Crudely stated: the energy that goes into a system must either come out or
accumulate. In the steady state, exactly the energy that goes into the engine must
come out, regardless of the efficiency of the engine. Clausius and Thomson both
saw in Carnot’s work that a perfect engine would conserve energy like any other
engine, but in addition, the quantity Q=T would also be conserved. In the perfect
(reversible) engine, Qin=Tin ¼ Qout=Tout. Not only is energy conserved (the first
law) but also the quantity Q=T is conserved in the ideal engine.
This astonishing fact leads us to the concept of entropy, S. Entropy was defined
then as now as S ¼ Qrev=T because of the remarkable discovery that the quantity
Q=T, entropy, is conserved in the ideal (reversible) engine. We can summarize the
work of Clausius, Thomson, and Carnot by saying that in an ideal engine, Sin ¼ Sout;
in any real engine, Sout  Sin. We now have a concise description of the most efficient
engine possible, Sin ¼ Sout. All real engines, regardless of their intended product,
produce entropy: SoutSin for any real engine.
This is the reason that the concept of reversibility is so important, and it leads
to the following definition:
Reversible process: A process in which there are no unbalanced driving forces. The
system proceeds from the initial state to the final state only because the driving forces
in that direction exceed forces opposing the change by an infinitesimal amount. In
short, no forces are wasted.

2.1. Master Concepts


Thermodynamics is built upon experience and experimentation. A few master con-
cepts have emerged, which may be stated as follows.
Postulate 1. There exists a form of energy, known as internal energy U, that is an
intrinsic property of the system, functionally related to pressure, temperature, and
the other measurable properties of the system. U is a property of the system, but
is not directly measurable.
The second master concept is simply a conservation equation, known as the
first law of thermodynamics:
Postulate 2. The total energy of any system and its surroundings is conserved.
A third postulate qualifies the first law by observing that not all forms of
energy have the same quality, or availability for use (the concept of entropy).
Postulate 3. There exists a property called entropy S, which is an intrinsic property
of the system that is functionally related to the measurable properties that character-
ize the system. For reversible processes, changes in this property can be calculated by

© 2005 by Marcel Dekker


5367-4 Flynn Ch02 R2 090804

Basic Principles 31

the equation
dS ¼ dQrev =T ð2:1Þ
where T is the absolute temperature of the system.
The first law of thermodynamics cannot be properly formulated without the
recognition of internal energy U as a property. Internal energy is a concept. Like-
wise, the second law of thermodynamics can have no complete description without
the recognition of entropy as a property. Entropy is also a concept, just as internal
energy is a concept, but is usually much harder to grasp.
Abstract quantities such as internal energy U and entropy S are essential to the
thermodynamic solution of problems. Accordingly, a necessary first step in thermo-
dynamic analysis is to translate the problem into the terminology of thermody-
namics. One of the real systems most often encountered in cryogenic engineering
is the one described by pressure, temperature, and specific volume. These systems
are called PVT systems and are composed of fluids, either liquids or gases or both.
For such systems, yet another postulate is required.
Postulate 4. The macroscopic properties of homogeneous PVT systems in
equilibrium states can be expressed as functions of pressure, temperature, and
composition only.
This postulate is the basis for all the thermodynamic equations that will follow
and has enormous utility. It is the basis for assuming the existence of an equation of
state relating P, V, and T. Note that it neglects the effects of electric, magnetic, and
gravitational fields and also neglects viscous shear and surface effects.
As mentioned, these postulates are primitive; they cannot be derived from
anything more basic.

2.2. The First Law of Thermodynamics


As stated above, the first law is a conservation equation: energy transferred to a
system will be conserved as changes in properties of the system or changes in
potentials of the system. That is, energy may be transferred or altered in form but
never created or destroyed. Energy that enters a system must either come out of
the system or accumulate (the checkbook principle):
Heat added to the system þ Work done on the system
¼ Changes in internal energy U
þ Changes in potential and kinetic energy
or
Q  W ¼ DU þ Dðpotential and kinetic energyÞ ð2:2Þ
In Eq. (2.2), the term Q  W represents all the energy a system has received from its
surroundings. This energy gained by the system is divided into two portions:
1. The increase in internal energy U within the system due to the composite
effect of changes in the configuration and motion of all of its ultimate
particles. This increment is represented by the term DU.
2. The increase in potential and kinetic energy of the entire system due
to changes in the position and motion of the system as a whole. This is

© 2005 by Marcel Dekker


5367-4 Flynn Ch02 R2 090804

32 Chapter 2

written as
mgDZ mDv2
DðPE + KEÞ ¼ þ ð2:3Þ
gc 2gc
where m is the mass of the system, DZ is the elevation of the system above
a reference plane, v is the velocity of the system, g is the local acceleration
of gravity, and gc is the gravitational constant.

Example 2.1. Water is flowing at a velocity of 100 ft=s through a pipe elevated 250 ft
above a reference plane. What are the kinetic energy (KE) and the potential energy
(PE) of 1 lbm of the water?
The work done on a body to increase its velocity leads to the kinetic energy
concept. From the first law, the work done on the body to accelerate it is trans-
formed into the kinetic energy of the body.
Work is defined as force times displacement, or
W ¼F l
and
dW ¼ F dl for a constant force F
The force may be obtained from Newton’s second law of motion,

ma

gc

where m is the mass of the body, a is the acceleration, and gc is a propor-


tionality constant called the gravitational constant. In the American engineering
system,
ðftÞðlbm Þ
gc ¼ 32:174
ðs2 Þðlbf Þ

Therefore,
ma
dW ¼ dl
gc

By definition, acceleration is
dv

dt

where v is the velocity of the body. Hence,


m dv
dW ¼ dl
gc dt

or
m dl
dW ¼ dv
gc dt

© 2005 by Marcel Dekker


5367-4 Flynn Ch02 R2 090804

Basic Principles 33

Since the definition of velocity is


dl

dt

the expression for work becomes


m
dW ¼ v dv
gc

This equation may now be integrated for a finite change in velocity from v1 to v2:
Z  
m v2 m v2 2 v1 2
W¼ v dv ¼ 
gc v 1 gc 2 2

or

mv2 2 mv1 2 Dmv2


W¼  ¼
2gc 2gc 2gc

The quantity mv2=2gc was defined as the kinetic energy by Lord Kelvin in 1856.
Thus,

mv2
KE ¼ ð2:4Þ
2gc

For the case given in the problem statement above,


    
ft 2 1 1 s2 lbf
Kinetic energy ¼ ð1 lbm Þ 100
s 2 32:174 ft lbm
¼ 155 ft lbf

Similarly, the work done on a body to change its elevation leads to the concept
of potential energy.
If a body of mass m is raised from an initial elevation z1 to a final elevation z2,
an upward force at least equal to the weight of the body must be exerted on it, and
this force must move through the distance z2  z1. Since the weight of the body is the
force of gravity on it, the minimum force required is given by Newton’s second law
of motion:
1 1
F¼ ma ¼ mg
gc gc
where g is the local acceleration of gravity and gc is as before. The minimum work
required to raise the body is the product of this force and the change in elevation:
g
W ¼ F ðz2  z1 Þ ¼ m ðz2  z1 Þ
gc
or
 
g g mzg
W ¼ mz2  mz1 ¼ D
gc gc gc

© 2005 by Marcel Dekker


5367-4 Flynn Ch02 R2 090804

34 Chapter 2

The work done on a body in elevating it is said to produce a change in its poten-
tial energy (PE), or
mzg
W ¼ DPE ¼ D
gc
Thus, potential energy is defined as
g
PE ¼ mz ð2:5Þ
gc
This term was first proposed in 1853 by the Scottish engineer William Rankine
(1820–1872).
For the problem statement above,
  
ft 1 s2 lbf
Potential energy ¼ ð1 lbm Þð250 ftÞ 32:174 2
s 32:174 ft lbm
¼ 250 ft lbf
Potential energy and kinetic energy are obvious mechanical terms. The energy
put into a system may be stored in elevation change or in velocity change as well as in
internal energy U by way of property changes.
The internal energy U represents energy stored in the system by changes in its
properties T, P, V, etc. DU is a composite effect of changes in the motion and
configuration of all of the ultimate particles of a system.
It is essential to note that the left-hand side of Eq. (2.2) describes a system–
surroundings interaction, while the right-hand side describes properties of the
system alone. This is a very valuable aspect of the first law equation, for it permits
the calculation of system–surroundings interactions (heat and work) from changes in
system properties only. Some of the consequences of this attribute of the first law
equation are discussed below.
As stated earlier, work is a system–surroundings interaction—energy in transit
between the system and surroundings, a precise kind of energy that crosses the
system boundary. Therefore, work must always have two components: an external
(surroundings) force and an internal (system) displacement. Work is always
Fexternal  Xsystem
where X is the system displacement.
For fluids, work is always
Pexternal  Vsystem
Only in the reversible case is Pexternal ¼ Pinternal. Then, and only then, may the P of
the system be substituted for Pexternal and equations of state be used to relate P
and V.
Because the left-hand side of Eq. (2.2) involves a system–surroundings interac-
tion, questions of path come into play (e.g., adiabatic, reversible, constant external
pressure, etc.). Indeed, the path must always be specified for calculations involving
the left-hand side of Eq. (2.2), i.e., heat and work. In thermodynamics, it is essential
to distinguish between quantities that depend on path and those that do not. Indeed,
we commonly make this distinction in ordinary affairs. For instance, the straight-line
distance between Denver and Los Angeles is a fixed quantity, a property. On the
other hand, other quantities for an auto-mobile trip between Denver and Los

© 2005 by Marcel Dekker


5367-4 Flynn Ch02 R2 090804

Basic Principles 35

Angeles—miles traveled, miles per gallon consumed, etc.—depend very much on the
path.
As stated, the right-hand side of Eq. (2.2) describes attributes of the system
only and contains properties and potentials of the system only. Thus, it can always
be evaluated from system properties or potentials alone without considering the
surroundings. The question of path is simply irrelevant, because there is no
system–surroundings interaction to be found on the right-hand side of the first law
statement.
It is a constant goal of thermodynamics to evaluate system–surroundings inter-
actions (e.g., heat and work) in terms of properties and potentials of the system
alone. The essence of the first law concept is (a) that it is a conservation equation
and (b) that a system–surroundings interaction can be expressed as a function of
system parameters only.
The first law is valid for all systems under all cases, as written in Eq. (2.2).
There are numerous special cases, and some of the more important will now be
considered.

2.2.1. Open System—the Flow Case


One special case of the first law of thermodynamics is the ‘‘open’’ system, or the flow
case. For a flow system such as the one shown in Fig. 2.5, the first law can be stated
as
g Dv2
Q  W ¼ DU þ DZ þ
gc 2gc

In this case the system is the element of fluid, and


W ¼ W s þ W1 þ W2
where Ws is the shaft work by pump, W1 the work done by surroundings on system
(element of fluid) pushing it past boundary 1 and W2 the work done by system on
surroundings as element of fluid emerges.
Since pressure is the force per unit area, F1 becomes
F1 ¼ P1 A1

Figure 2.5 Open thermodynamic system—the flow case.

© 2005 by Marcel Dekker


5367-4 Flynn Ch02 R2 090804

36 Chapter 2

and the distance l1 through which F1 acts is


V1
l1 ¼
A1
where V1 is the specific volume of element fluid and A1 the cross-sectional area
of pipe.
Therefore, W1 ¼ F1P1 becomes
 
V1
W1 ¼ P1 A1 ¼ P1 V1
A1
Likewise,
 
V2
W2 ¼ P2 A2 ¼ P2 V2
A2
and the total work, W, is
W ¼ Ws þ P2 V2  P1 V1
The first law, as written above, becomes for this special case
g Dv2
Q  ðWs þ P2 V2  P1 V1 Þ ¼ DU þ DZ þ
gc 2gc
g Dv2
Q  Ws ¼ DU ¼ DZ þ þ P2 V2  P1 V1
gc 2gc
Dv2 g
¼ U2 þ P2 U2  U1  P1 V1 þ þ DZ
2gc gc
which defines the enthalpy H as
H  U þ PV ð2:6Þ
Then,

Dv2 g
Q  Ws ¼ H 2  H 1 þ þ DZ
2gc gc
or

Dv2 g
Q  Ws ¼ DH þ þ DZ
2gc gc
If velocity and elevation changes are negligible, then Q  Ws ¼ DH. This is the
flow system analog of the first law for closed systems.
Consideration of other special cases of the first law leads to other auxiliary
functions, discussed below.

2.2.2. The Constant-Volume Case


The first law, Q  W ¼ DU, can always be written in differential form because no
path is involved in differential notation:
dQ  dW ¼ dU

© 2005 by Marcel Dekker


5367-4 Flynn Ch02 R2 090804

Basic Principles 37

But dW ¼ PextdV, and since dV ¼ 0 for constant volume,


dQ ¼ ðdUÞV
Differentiating with respect to temperature suggests the function
   
dQ dU
¼  CV
dT V dT V
which is the definition of heat capacity at constant volume.

2.2.3. The Constant-Pressure Case


The first law in differential form can be combined with the expression for work under
constant pressure, dW ¼ P0 dV, to yield
dQ  dW ¼ dU
Thus,
dQ ¼ P0 dV þ dU
where P0 ¼ the constant system pressure. Rearranging gives

QP ¼ P0 ðV2  V1 Þ þ DU
¼ P0 V2  P0 V1 þ U2  U1
¼ ðP0 V2 þ U2 Þ  ðP0 V1 þ U1 Þ
and since P0 ¼ P2 ¼ P1 ¼ constant, this may be rewritten as
QP ¼ ðP2 V2 þ U2 Þ  ðP1 V1 ¼ U1 Þ
which, from the definition of enthalpy, H  U þ PV
QP ¼ H 2  H 1 ¼ DH
Thus, at constant pressure,
Q ¼ ðDHÞP
Differentiating this expression with respect to temperature suggests the function
   
dQ dH
¼  CP
dT P dT P
which is the definition of heat capacity at constant pressure.

2.3. The Second Law of Thermodynamics


Another master concept, the second law, comes about from observing that Q and W
do not have the same ‘‘quality’’. For instance,
1. W is highly ordered and directional.
2. Q is less versatile.
3. W can always be completely converted to Q regardless of the temperature,
but not vice versa.
4. Q can flow only from Thot to Tcold.

© 2005 by Marcel Dekker


5367-4 Flynn Ch02 R2 090804

38 Chapter 2

Q has a quality as well as a quantity, and that quality depends on the temperature.
Also, the effect of adding Q to a system is always to change the properties of
the system, so there must be a property change that describes the addition of Q.
We postulate a property in thermodynamics to describe these observations that
meets the following qualitative requirements:
1. The addition of Q causes a property change in the system.
2. The way (path) in which Q is added fixes the magnitude of the property
change.
3. Adding Q at a lower temperature causes a greater change than adding the
same amount of Q at a higher temperature.

Thus, we invent a property of the system to fit these three criteria, namely


entropy, S (see Eq. (2.1)):
Qrev

T
or

Qrev ¼ TDS

Since Q is not a property, its path must be specified, and Qrev is chosen to complete
the definition.
Thus, we have developed qualitatively Postulate 3, which was stated earlier:
there exists a property called entropy S, which is an intrinsic property of a system
that is functionally related to the measurable coordinates that characterize the
system. For a reversible process, changes in this property are given by
dQrev
dS ¼
T
The definition of entropy, Qrev ¼ TDS, can always be written in differential
form because neither reversibility nor any other path has any significance in differ-
ential notation:
dQ ¼ T dS

and hence the first law,

U ¼QW

becomes in differential form, for all cases and for all substances,

dU ¼ T dS  P dV ð2:7Þ

This is a universal relationship, just like the first law. One can always write the first
law in this differential form for all cases. Although derived for a reversible process,
Eq. (2.7) relates properties only and is valid for any change between equilibrium
states in a closed system. The question of reversibility=irreversibility has no signifi-
cance to Eq. (2.7) because (a) the concept of path has no bearing on any equation
written in differential form and (b) the equation dU ¼ T dS  P dV contains proper-
ties of the system only. Path is a concept relevant to system–surroundings interac-
tions, and ‘‘surroundings’’ are completely absent from this description of the
system alone.

© 2005 by Marcel Dekker


5367-4 Flynn Ch02 R2 090804

Basic Principles 39

Entropy is a property of the system that is dependent on the configuration of


all the ultimate particles that make up the system. A system in which all the ultimate
particles are arranged in the most orderly manner possible has an entropy of zero. As
the arrangement within the system is changed from an ordered to a more disordered
or more random state, the entropy of the system increases.
The addition of heat to a system causes an increase in scattering and disorder
among all the particles within the system, in other words, an increase in entropy.
If heat is added during a reversible process, this increase in entropy is related to
the heat added by
Z
Qrev ¼ T dS

If heat is transferred to the system during an irreversible process, then


Z
T dS > Q

or
Z
T dS ¼ Q þ lost work

The lost work is energy that might have done useful work but is dissipated instead,
causing unnecessary randomness and disorder within the system.
The second law can be stated formally as follows.
Second law of thermodynamics. The entropy change of any system and its surround-
ings, considered together, is positive and approaches zero for any process that
approaches reversibility.
The second law of thermodynamics is a conservation law only for reversible
processes, which are unknown in nature. All natural processes result in an increase
in total entropy. The mathematical expression of the second law is simply
DStotal  0
where the label ‘‘total’’ indicates that both the system and its surroundings are
included. The equality applies only to the limiting case of a reversible process.

2.4. Clausius Inequality and Entropy


In 1824, Nicholas Leonard Sadi Carnot proposed a heat power cycle composed of
the following reversible processes shown in Fig. 2.6.
1. Isothermal reversible expansion and absorption of heat Q1 at temperature
T1 (a–b).
2. Isentropic (adiabatic, reversible) expansion to T2 (b–c).
3. Isothermal reversible compression and rejection of heat Q2 at T2 (c–d).
4. Isentropic compression back to temperature T1 (d–a).

Since the energy change of the system for the complete cycle must be zero,
DUcycle ¼ 0 ¼ Q1  Q2  W

© 2005 by Marcel Dekker


5367-4 Flynn Ch02 R2 090804

40 Chapter 2

Figure 2.6 Carnot’s reversible heat power cycle.

or

Q1  Q2 ¼ W
where W is the net work, which is the difference between the work produced in the
expansion process (b–c) and the work required in the compression process (d–a).
Also, the absolute values of Q1 and Q2 are used, temporarily abandoning the usual
sign convention.
The efficiency of this reversible cycle is
W Q1  Q2
Z¼ ¼
Q1 Q1
since Q1 ¼ T1DS and Q2 ¼ T2DS
The efficiency of the Carnot engine can be expressed in terms of the tempera-
ture of the heat reservoirs alone:
T1  T2

T1
Carnot stated (and proved) the following principles, which are known today as
the Carnot principles:
1. No engine can be more efficient than a reversible engine operating between
the same high-temperature and low-temperature reservoirs. Here the term
‘‘heat reservoir’’ is taken to mean either a heat source or a heat receiver or
sink.
2. The efficiencies of all reversible engines operating between the same
constant-temperature reservoirs are the same.
3. The efficiency of a reversible engine depends only on the temperatures of
the heat source and heat receiver.

The two equations for the efficiency of the Carnot cycle can be combined to
yield
 
W Q1  Q2 T1  T2 T2
Z¼ ¼ ¼ ¼ 1
Q1 Q1 T1 T1
or
 
T2
W ¼ Q1 1 
T1

© 2005 by Marcel Dekker


5367-4 Flynn Ch02 R2 090804

Basic Principles 41

But

W ¼ Q1  Q2

Therefore,
 
T2
Q1  Q2 ¼ Q1 1 
T1

which reduces to

Q1 Q2
¼
T1 T2
or S1 ¼ S2, which is to repeat that entropy is conserved in a reversible engine.
In 1848, Lord Kelvin (William Thomson of Glasgow) derived a temperature
scale based on the above equation, independent of the thermodynamic substance
employed in the cycle. He defined the scale such that
T 1 Q1
¼
T 2 Q2

where Q1 is the heat received by a Carnot engine from a source at temperature T1


and Q2 is the heat rejected to a receiver at temperature T2. One advantage of the
absolute scale is the absence of negative temperatures.
Clausius noted that the Kelvin equation could be rearranged to give
Q1 Q2 Q1 Q2
¼ or þ ¼0
T1 T2 T1 T2
Here, the usual sign convention is used and the minus sign with Q2 denotes that
heat is leaving the system. This equation indicates that the change in the quantity
Q=T around a reversible or Carnot cycle is zero. Clausius extended this relationship
to embrace all cyclic processes by stating that the cyclic integral of dQ=T is less than
zero or in the limit equal to zero; thus,
I
dQ
0
T
This equation is now known as the inequality of Clausius.
It can be proved that the cyclic integral of dQ=T is equal to zero for all rever-
sible cycles and less than zero for all irreversible cycles. Since the summation of the
quantity dQ=T for a reversible cycle is equal to zero, it follows that the value of the
integral of dQ=T is the same for any reversible process between states 1 and 2 of a sys-
tem and thus is a property of the system. It is this proof that gives credence to Postu-
late 3, that there is a property called entropy that is defined as the ratio of heat
transferred during a reversible process to the absolute temperature of the system,
or mathematically,
 
dQ
ds ¼
T rev

Entropy, since it is a property, is especially useful as one of the coordinates when


representing a reversible process graphically.

© 2005 by Marcel Dekker


5367-4 Flynn Ch02 R2 090804

42 Chapter 2

2.5. The Maxwell Relations


We are now in a position to construct a complete thermodynamic property network
and develop the Maxwell relations. The purpose is to develop a set of property func-
tions (also known as state functions or thermodynamic functions) that will enable us
to calculate any thermodynamic property change (e.g., DS, DH, DU) for any
substance for any process.
By definition, H ¼ U þ PV and
dH ¼ dU þ P dV þ V dP
or
dU ¼ dH ¼ P dV  V dP
This form of U may be substituted into the first law in differential form:
dU ¼ T dS þ P dV ð2:8Þ
to yield
dH ¼ T dS þ V dP ð2:9Þ
It can easily be shown that for a reversible flow process,
Wmax ¼ H  Qrev
or
dW ¼ dH  T dS
This form of the work available from a reversible flow process suggests the function
G ¼ H  TS ð2:10Þ
which is the definition of Gibbs free energy. This can likewise be substituted into the
differential form of the first law to yield
dG ¼ S dT þ V dP ð2:11Þ
Similarly, the maximum work for a closed reversible process is
Wmax ¼ DU  Qrev
or
dW ¼ dU  T dS
which suggests another new function,
A ¼ U  TS ð2:12Þ
which is known as the Helmholtz free energy. If this is likewise substituted into the
first law in differential form, we have
dA ¼ S dT  P dV ð2:13Þ
We have thus defined the state functions, functions of the properties of a system
that, under various conditions, can relate changes in the properties of the system to
the heat and work transferred between it and the surroundings. Since these functions
or potentials are functions only of the properties of the system, they can be expressed

© 2005 by Marcel Dekker


5367-4 Flynn Ch02 R2 090804

Basic Principles 43

as exact differentials. Thus, changes in these functions depend only on the initial and
final states of the system and not on the method or path by which the change takes
place. These functions are
U ¼QW Internal energy
H ¼ U þ PV Enthalpy
G ¼ H  TS Gibbs free energy
A ¼ U  TS Helmholtz free energy
The four state functions written in differential form are
dE ¼ T dS  P dV ðEq: 2:8Þ
dH ¼ T dS þ V dP ðEq: 2:9Þ
dG ¼ S dT þ V dP ðEq: 2:11Þ
dA ¼ S dT  P dV ðEq: 2:13Þ
It would be imprecise to stop with the impression that these four state func-
tions are derived from considering the maximum possible work connected with a
reversible process. Such a consideration only suggests the function. New thermody-
namic functions such as H, G, and A cannot in general be defined by the random
combination of variables. As a minimum, dimensional consistency is required.
Instead, a standard mathematical method exists for the systematic definition of
the desired functions; this is the Legendre transformation. The functions for G, H,
and A are in fact Legendre transformations of the fundamental property relation
of a closed PVT system, dU ¼ TdS  PdV.
A consequence of the Legendre transformation is that the resulting expressions
for G, H, and A are exact differential equations with the following special property:
If z ¼ f(x,y) and dz ¼ Mdx þ Ndy, then the property of exactness is that
   
@M @N
¼
@y x @x y

Applying this property of exact differential equations to the four equations above
yields the following Maxwell relations:
   
@T @P
¼ ð2:14Þ
@V S @S V
   
@T @V
¼ ð2:15Þ
@P S @S P
   
@S @V
¼ ð2:16Þ
@P T @T P
   
@S @P
¼ ð2:17Þ
@V T @T V

The last two are among the most useful, because they relate entropy changes to an
equation of state.
The development of many of these relations is shown in Fig. 2.7, as proposed
by Dr. A. J. Kidnay of the Colorado School of Mines.

© 2005 by Marcel Dekker


44
Chapter 2
Figure 2.7 The thermodynamic network.

© 2005 by Marcel Dekker


5367-4 Flynn Ch02 R2 090804

Basic Principles 45

2.6. Equations of State


It is evident from the preceding discussion that thermodynamics provides a multitude of
equations inter-relating the properties of substances. Given appropriate data, thermody-
namics allows the development of a complete set of thermodynamic property values from
which one can subsequently calculate the heat and work effects of various processes.
PVT data are among the most important and are concisely given by equations of state.
The simplest equation of state is the ideal gas law:
PV
¼1 ð2:18Þ
RT
The ideal gas law is useful for calculating limiting, low-pressure, cases; for defining
the ideal gas thermodynamic temperature scale; and for calculating so-called residual
properties; but it fails to predict the existence of the liquid state.
The ideal gas law, with all of its shortcomings, is nonetheless very useful and is
found at the core of all other equations of state. It is fun to develop from mechanics.
Doing so shows us that temperature is a direct measure of kinetic energy and that
pressure is kinetic energy per unit volume.

2.6.1. Deriving the Ideal Gas Law


The requisites for an ideal gas are that the molecules occupy zero volume (or negli-
gible volume compared to the total available) and that there be no interactions
(forces) between molecules.
Consider n molecules of an ideal gas confined in a volume V t. We want to find
the pressure, i.e., the force per unit area, exerted on a confining wall. Force is the
time rate of change of momentum, so we want to find that rate

The molecules are in random motion. Their velocity, a vector, can be resolved into
three components. Take one component normal to the wall. Since the motion is ran-
dom, on the average, one-third of the velocities are normal to the wall and half of
those, one-sixth of the total, are moving toward the wall at any time (the other half
are moving away from the wall).
The total number of molecules in the small box shown in the figure is the number
per unit volume, n=V t, times the volume A dL. One-sixth of that number is moving
toward the wall and will strike the wall in the time dt that it takes to travel the distance
dL at the average velocity, u ¼ dL=dt. Thus, the rate at which molecules strike the wall is
 
1 n  1 1 n   u  1 n 
A dL ¼ A dL ¼ Au
6 Vt dt 6 Vt dL 6 Vt

© 2005 by Marcel Dekker


5367-4 Flynn Ch02 R2 090804

46 Chapter 2

When molecules strike the wall, their velocity is reversed. So the total change in velocity
is 2u. The rate of change of momentum is the rate of striking the wall times the mass per
molecule (i.e., molecular weight times the change in velocity, 2u). Thus,
 
1 n  1 Mn
F¼ AuðMÞð2uÞ ¼ Au2
6 Vt 3 Vt

and the pressure is


 
F 1 Mn 2
P¼ ¼ u
A 3 Vt

The ideal gas law is P ¼ (n=V t) RT; thus,

1 Mu2 3
Mu2 ¼ RT or ¼ RT
3 2 2
where Mu2=2 is the average kinetic energy per mole. Thus, we have shown that for an
ideal gas the kinetic energy is directly proportional to the absolute temperature. Or, put
another way, temperature is a direct measure of kinetic energy. Notice also that
pressure is equal to kinetic energy per unit volume.
This short derivation also provides some insight into the universal gas
constant R. R is the chemist’s version of Boltzmann’s constant k. R is in units per
mole, and k is in units per molecule. Both R and k are scaling factors between tem-
perature and energy. For physicists, E ¼ kT. For chemists, it is PV ¼ RT, where the
product PV is energy. Several values of R are given in Table 2.1.
We have already defined the heat capacities Cv and Cp. The ideal gas derivation
can also breathe some life into these two important quantities. Heat capacity
describes the ability of some substance to ‘‘soak up’’ energy.
Consider the molecules of a monatomic ideal gas. The only way a monatomic
gas can soak up energy is to increase in kinetic energy. That is, the only way a mona-
tomic ideal gas can ‘‘contain’’ energy (short of energies large enough to affect elec-
tronic or nuclear changes) is by the motion of its molecules—translational kinetic
energy. Thus, if energy is added to a box of such molecules at constant volume,
the change in their internal energy will be

dU ¼ dðkinetic energyÞ ¼ dðð3=2ÞRTÞ ¼ ð3=2ÞR dT

At constant volume, dU ¼ CV dT, therefore CV ¼ (3=2) R for a monatomic ideal gas.


If energy is added at constant pressure, the volume will change. Work is done when
the volume changes. The work is P dV, due to pushing back the surroundings at the
same pressure, P, as the system. The total energy added is the change in internal
energy plus the work (dU þ P dV). This quantity equals dH (at constant P), and
dH ¼ CP dT (at constant P). So,

dH ¼ CP dT ¼ dU þ P dV ¼ CV dT þ R dT ¼ ðCV þ RÞ dT

Thus, CP ¼ CV þ R. This is true for any ideal gas (monatomic or not, because the
work expression is always the same).
For polyatomic molecules, the internal energy includes the translational kinetic
energy, but the molecules also can ‘‘soak up’’ energy in the form of rotation and
vibration about their bonds. Polyatomic molecules thus have greater CV (and CP)

© 2005 by Marcel Dekker


5367-4 Flynn Ch02 R2 090804

Basic Principles 47

Table 2.1 Values of the Gas Constant R for 1 mol of Ideal Gas

Temp. unit Pressure unit Volume unit Energy unit R per gram mol

K Calorie 1.9872
K atm cm3 82.057
K atm L 0.082054
K mmHg L 62.361
K bar L 0.08314
K kg=cm2 L 0.08478

R per lb mol

R Btu (IT) 1.986

R hp hr 0.0007805

R kW hr 0.0005819

R atm cu ft 0.7302

R in. Hg cu ft 21.85

R mmHg cu ft 555.0

R psia cu ft 10.73
psfa cu ft ft lb 1545.0
K atm cu ft 1.314
K mmHg cu ft 998.9

Selected values from API Research Project 44:


Ice point ¼ 0 C ¼ 273.16 K ¼ 491.69 R.
Gram molar volume of ideal gas at 1 atm and 0 C ¼ 22,414.6 cm3 ¼ 22.4140 L.
Pound molar volume of ideal gas at 1 atm and 32 F ¼ 359.05 cu ft.
1 atm ¼ 760 mmHg ¼ 29.921 in. Hg ¼ 14.696 psia.
1 hp ¼ 550 ft lb=s ¼ 745.575 W (IT).
1 cal ¼ 41833 J (IT).
1.8 Btu=lb ¼ 1.0 IT cal=g ¼ 1.000657 cal=g.

than do monatomic molecules. As a first approximation, CV increases by R for each


additional atom in the molecule. This is a consequence of the principle of equipar-
tion of energy: the value of CV increases by (1=2)R for each additional mode by
which energy can be soaked up (three translational modes plus one rotational and
one vibrational per bond formed). This holds well for diatomic molecules
[CV ¼ (5=2)R, CP ¼ (7=2)R], but as more atoms are added it does not work so well.
The vibrational modes and rotational modes are not completely independent in more
complex molecules.
The molecules of real gases do occupy a volume of their own, and there is an
attraction among them, the intermolecular force. The equation of state for a real gas
will therefore contain the ideal gas law as a first approximation and add other terms
to account for the volume of the molecules and the potential between the molecules.
These other terms express the nonideality of real gases. We need these terms, not
only for accuracy but also to explain why some gases cool as they expand and
others do not. Without the nonideality expressed by these terms, there would be
no change in temperature upon expansion. Ideal gases have a zero Joule–Thomson
coefficient.
It is easy to find the ideal gas law in the virial equation of state.
The virial equation of state is extremely powerful and can be written in two
forms.

© 2005 by Marcel Dekker


5367-4 Flynn Ch02 R2 090804

48 Chapter 2

1. As an expansion in volume (V) or density (r):


PV B C
¼ 1 þ þ 2 þ  ð2:19Þ
RT V V
or

PV
¼ 1 þ Br þ Cr2 þ    ð2:20Þ
RT
2. As an expansion in pressure:
PV
¼ 1 þ B 0 P þ C 0 P2 þ    ð2:21Þ
RT

The virial coefficients (B, C, . . . ) can be calculated from first principles (two-body
interactions, three-body interactions, etc.) and are functions of temperature only.
Generalized or corresponding state forms of the equation of state are also useful:

PV
¼ Z; Z ¼ ZðTr ; Pr Þ ð2:22Þ
RT

where Z is the compressibility factor and Tr and Pr are the reduced temperature and
pressure, e.g., Tr ¼ T=Tc. T is the system temperature, and Tc is the critical tempera-
ture of the system component.
Another useful generalized equation of state is the three-parameter corre-
sponding states form employing the acentric factor o,

Z ¼ f ðTr ; Pr ; oÞ ð2:23Þ

where o is a tabulated parameter for a specific substance.


Several of the more common equations of state are shown in Table 2.2.
The values of DS, DH, DE, etc., can now be calculated for any process for any
substance according to the following scheme.
Step 1. Choose a pair of coordinates from P, T, and V, for example, P–T, T–V, or
P–V. The problem statement is useful; use what is given. If the process is isothermal
or isobaric, choose T (or P) as one of the pair, because it is constant.
The P–T pair is commonly chosen (but any two will do for homogeneous sys-
tems of a pure component), as P and T are the most commonly measured variables.
Step 2. Write the total differential in terms of the pair chosen, e.g.,

S ¼ f ðP; TÞ
   
@S @S ð2:24Þ
dS ¼ dT þ dP
@T P @P T

Step 3. Evaluate the partial differential coefficients from the basic Maxwell relations.
For example, evaluate the first coefficient, (@S=@T)P. Because this coefficient
involves an entropy term (S) that changes with T at constant P, it is likely that H
will be involved.
From the definition of dH,

dH ¼ T dS þ V dP

© 2005 by Marcel Dekker


5367-4 Flynn Ch02 R2 090804

Basic Principles 49

Table 2.2 Equations of State (for 1 mol)

Van der Waals:


! Beattie–Bridgeman:
a b  bÞ ¼ RT b ¼ RT þ b þ g þ d
p þ 2 ðV pV
b
V b V
V b2 V
b3
Lorenz:
RT b a Rc
p ¼ 2 ðV þ bÞ  2 b ¼ RTBo  Ao 
b
V Vb T2
RBo c
Dieterici: g ¼ RTBo b þ aAo 
T2
RT a=VbRT RBo bc
p¼ e d¼
b b
V T2
Berthelot: Benedict–Webb–Rubin:
RT a b ¼ RT þ b þ s þ Z þ o
p¼  pV
b
V  b TV b2 b V
V b2 V b4 V b5
Redlich–Kwong:
" #
a b  bÞ ¼ RT Co
pþ ðV b ¼ RTBo  A0 
b ðV
T 1=2 V b þ bÞ T2
 
R2 Tc2:5
a ¼ 0:4278 s ¼ bRT  a þ Tc2 exp  g2
pc   b
V
RTc g
b ¼ 0:0867 Z ¼ cg exp  2
pc b
V
o ¼ aa
Onnes: ! Peng–Robinson:
b ¼ RT 1 þ B C RT aa
pV þ 2 þ  p¼ 
b V
V b b b V
V b ðVb þ bÞ þ bðV
b  bÞ
Holborn:  2 2
b ¼ RTð1 þ B0 p þ C 0 p2 þ   Þ R Tc
pV a ¼ 0:45724
pc
 
RTc
b ¼ 0:07780
pc
a ¼ ½1 þ kð1  Tr1=2 Þ2
k ¼ 0.37464 þ 1.5422o  0.26992o2
o ¼ acentric factor

we have
   
@H @S
¼T þ0
@T P @T P
Since at constant pressure dP ¼ 0, and since
 
@H
 CP
@T P
 
@S
CP ¼ T
@T P
therefore,
 
@S 1
¼ CP ð2:25Þ
@T P T
So far, there are no restrictions on this equation.

© 2005 by Marcel Dekker


5367-4 Flynn Ch02 R2 090804

50 Chapter 2

Step 4. Now evaluate the second coefficient, (@S=@P)T. From the relationship
dG ¼ S dT þ V dP
and the property of exactness described before, we have
   
@S @V
¼ ð2:26Þ
@P T @T P
Step 5. The total differential, dS, is now complete as
 
CP @V
dS ¼ dT  dP ð2:27Þ
T @T P

There are still no restrictions on this equation.


Step 6. Now evaluate (@V=@T)P from an equation of state or numerically. If we
choose PV ¼ RT as the arbitrarily selected equation of state, then
 
@V R
V ¼ RT=P and ¼
@T P P

and
dT dP
dS ¼ CP R
T P
We now have the total derivative of S as a function of properties only for any
process. The integration to yield DS is straightforward. This equation is now restricted
to the ideal gas state, since that was the equation of state chosen to evaluate @V=@TP.
This is a perfectly general method for any thermodynamic function: DS, DH,
DU, DG, etc. Since such equations contain only properties of the system, they are
independent of path and can be tabulated for various fluids. Such tabulations form
the basis for the various thermodynamic charts: T–S, H–S, etc.
Example 2.2. Calculate internal energy (U) and entropy (S) as a function of T
and V.
It is sometimes more convenient to take T and V as independent variables
rather than T and P. The method described above is perfectly general and can be
applied here. Writing U and S as total differentials:
   
@U @U
dU ¼ dT þ dV ð2:28Þ
@T V @V T
   
@S @S
dS ¼ dT þ dV ð2:29Þ
@T V @V T

Recall that
 
@U
CV ¼ ð2:30Þ
@T V

Two relations follow immediately from differentiating Eq. (2.8):


   
@U @S
¼T ð2:31Þ
@T V @T V

© 2005 by Marcel Dekker


5367-4 Flynn Ch02 R2 090804

Basic Principles 51

and
   
@U @S
¼T P ð2:32Þ
@V T @V T

As a result of Eq. (2.30), Eq. (2.31) becomes


 
@S CV
¼ ð2:33Þ
@T V T
and as a result of Eq. (2.17), Eq. (2.32) becomes
   
@U @P
¼T P ð2:34Þ
@V V @T V
Combining Eqs. (2.28), (2.30), and (2.34) gives
   
@P
dU ¼ CV dT þ T P dV ð2:35Þ
@T V
and combining Eqs. (2.29), (2.17), and (2.33) gives
 
CV @P
dS ¼ dT þ dV ð2:36Þ
T @T V
Equations (2.35) and (2.36) are general equations that express the internal energy
and entropy of homogeneous fluids at constant composition as functions of tempera-
ture and molar volume. The coefficients of dT and dV are expressed in terms of
measurable quantities.

2.6.2. Useful Thermodynamic Relations


With these basic definitions and concepts we can now obtain a whole host of useful
thermodynamic relations. Table 2.3 lists these relationships as functions of the pro-
cess for both real systems and ideal gas systems. A distinction is also made, where
necessary, between nonflow and flow systems, which are sometimes designated as
closed and open systems, respectively.

2.7. Thermodynamic Analysis of Low-Temperature Systems


The fundamental tools for a low-temperature analysis are the first and second laws
of thermodynamics. Assuming that the kinetic and potential energy terms in the first
law can be neglected (frequently a valid assumption), the first law for a steady-state
process such as a cryogenic refrigerator can be simplified to
DH ¼ Q  Ws ð2:37Þ
where DH is the summation of the enthalpy differences of all the fluids entering and
leaving the system (or piece of equipment) being analyzed, Q is the summation of all
heat exchanges between the system and its surroundings, and Ws is the net shaft
work. There can be as many independent first law analyses as there are independent
systems or pieces of equipment in the refrigeration and liquefaction systems. These
can involve numerous unknowns, some of which may have to be fixed as parameters.

© 2005 by Marcel Dekker


5367-4 Flynn Ch02 R2 090804

52 Chapter 2

Table 2.3 Useful Thermodynamic Relationships

Process W Q DU DH

Isobaric (constant P)
Real P(V2  V1) nCP(T2  T1) QW nCP(T2  T1)
Ideal gas nR(T2  T1) nCV(T2  T1)
Isometric (constant V)
Real 0 nCV(T2  T1) Q DU þ D(PV)
Ideal gas nCV(T2  T1) nCP(T2  T1)
Isothermal (constant T) Q  DU DU þ W QW DU þ D(PV)
Real nRT ln(V2=V1) nRT ln(P1=P2) 0 0
Ideal gas nRT ln(P1=P2) nRT ln(V2=V1)
Isentropic (constant S)
Real (nonflow) DU 0 W DU þ D(PV)
Ideal gas (nonflow) (P2V2  P1V1)=(1  g) nCV(T2  T1) nCP(T2  T1)
nCV(T2  T1)
Real (flow) DH 0 DH  D(PV) W
Ideal (flow) g(P2V2  P1V1)=(1  g)
Polytropic
Real (nonflow) Q  DU DU þ W QW DU þ D(PV)
(P2V2  P1V1)=(1  n)
Ideal (nonflow) Q  nCV(T2  T1) nCV(T2  T1) nCP(T2  T1)
Real (flow) Q  DH DH þ W DH  D(PV) QW
Ideal (flow) n(P2V2  P1V1)=(1  n)
Cyclic Qnet Wnet 0 0

The choice of these parameters will generally be made on the basis of experience or
convenience.
Application of the first law to the ideal work of isothermal compression in a
refrigerator cycle with appropriate substitution for the heat summation reduces to
Ws =m
_ ¼ T1 ðS1  S2 Þ  ðH1  H2 Þ ð2:38Þ

where ṁ is the mass rate of flow through the compressor, S is the specific entropy, H
is the specific enthalpy, and the subscripts 1 and 2 refer to the inlet and exit streams,
respectively, of the isothermal compressor. In a real system, an overall compressor
efficiency factor will have to be incorporated in this equation to obtain meaningful
results.
Evaluation of the refrigeration duty in a refrigerator cycle is dependent on the
process chosen. For example, if the refrigeration process is isothermal, the refrigera-
tion duty can be evaluated using the relation
Q¼m
_ TðS2  S1 Þ ð2:39Þ

where T is the temperature at which the refrigeration takes place and the subscripts 1
and 2 refer to the inlet and exit streams of the process. On the other hand, if the
refrigeration process involves only the vaporization of the saturated liquid refriger-
ant at constant pressure, evaluation of the duty reduces to

Q¼m
_ hfg ð2:40Þ

where hfg is the specific heat of vaporization of the refrigerant. If the process is
performed at constant pressure with the absorption of only sensible heat by the

© 2005 by Marcel Dekker


5367-4 Flynn Ch02 R2 090804

Basic Principles 53

refrigerant (e.g., warming a cold gas), the duty can be evaluated from
_ ðH2  H1 Þ
Q¼m ð2:41Þ
Similar first law balances could be written for other units of equipment used at low
temperatures.

2.8. Joule–Thomson Coefficient


The slope at any point of an isenthalpic curve on a temperature–pressure diagram is
called the Joule–Thomson (or J–T) coefficient. The J–T coefficient is usually denoted
by m and is given by
 
@T

@P H
The locus of all points at which the J–T coefficient is zero, i.e., the locus of the
maxima of the isenthalpic curves, is known as the inversion curve and is shown as a
dotted curve in Fig. 2.8. The region to the left and inside the inversion curve, where
the J–T coefficient is positive, is the region of cooling; the region outside, where the
J–T coefficient is negative, is the region of heating.
Hydrogen, helium, and neon have negative J–T coefficients at ambient tem-
perature. Consequently, when used as refrigerants in a throttling process, they must
be cooled either by a separate precoolant fluid or by a work-producing expansion to
a temperature below which the J–T coefficient is positive. Only then will throttling
cause a further cooling rather than heating.
Nitrogen, methane, and certain other fluids, on the other hand, have positive
J–T co-efficients at ambient temperatures and hence produce cooling when expanded
across a valve. Accordingly, any of these fluids can be used directly as the refrigerant
in a throttling process without the necessity of a precooling step or expansion
through a work-extracting device. Maximum inversion temperatures for some of
the more common cyrogens are given in Table 2.4.

Figure 2.8 Joule–Thomson inversion curve.

© 2005 by Marcel Dekker


5367-4 Flynn Ch02 R2 090804

54 Chapter 2

Table 2.4 Maximum Inversion Temperature

Maximum inversion temperature



Fluid K R

Oxygen 761 1370


Argon 722 1300
Nitrogen 622 1120
Air 603 1085
Neon 250 450
Hydrogen 202 364
Helium 40 72

Example 2.3. Show that the Joule–Thomson coefficient for a perfect gas is always
zero.
The Joule–Thomson coefficient has been defined as
 
@T
m¼ ð2:42Þ
@P H

From calculus it can be shown that


    
@T @T @H
mJT ¼ ¼ ð2:43Þ
@P H @H P @P T

The expression (@H=@T)P has previously been identified as the heat capacity, Cp. The
expression (@H=@P)T can be found by differentiating Eq. (2.9 ) and combining the
result with Eq. (2.16 ) to yield
    
@H @V
¼ V T
@T P @T P

The total differential for H,


   
@H @H
dH ¼ dT þ dP
@T P @P T

becomes
  
@V
dH ¼ CP dT þ V  T dP ð2:44Þ
@T P
Setting dH ¼ 0 (constant enthalpy), Eq. (2.45 ) gives the Joule–Thomson coeffi-
cient in terms of measurable thermodynamic properties as
   
1 @V
mJT ¼ T V ð2:45Þ
CP @T P
For a perfect gas, V ¼ RT=P, and
 
@V R V
¼ ¼
@T P P T

© 2005 by Marcel Dekker


5367-4 Flynn Ch02 R2 090804

Basic Principles 55

Therefore, for a perfect gas,


   
1 V
mJT ¼ T V ¼0 ð2:46Þ
CP T
A perfect gas would not experience any temperature change upon expansion.

3. HEAT TRANSFER

All cryogenic systems operate, that is to say complete their required tasks, at very
low temperatures. This fact requires a reasonably strong understanding of heat
transfer and all associated technologies. As stated earlier, the best that one can hope
for in some cases (e.g., cryogenic fluid storage) is to severely limit the energy transfer
across boundaries. Since heat always flows from hot to cold and cannot be stopped,
how it is controlled is the design aspect relating to heat transfer. In other cases (like
gasification of cryogenic fluids) the ability to maximize the amount of heat intro-
duced into the fluid is the desired design state. Due to the nature of the temperature
difference between the everyday world we live in and the extremes of operating ther-
mal conditions of cryogenic systems, a number of unique problems present them-
selves. These include, but not limited to: (1) variability of material properties, (2)
thermal insulation, (3) fluid property differences for near-critical-point convection,
(4) effects of thermal radiation, (5) equipment design characteristics and operation,
(6) selection of instrumentation and control sensors, (7) transfer of cryogenic fluids,
and (8) disposal and=or control of cryogenic boil-off. Each of these problems will be
addressed in the remaining chapters of this book. Heat transfer at low temperatures
is governed by the same three mechanisms present at ambient and elevated tempera-
tures: conduction, convection, and radiation. It is not surprising, therefore, that all
the general equations are appropriate for low-temperature applications as long as
they are adjusted for the property changes in both materials and fluids.

3.1 Conductive Heat Transfer


Another method of determining the heat transfer rate is in the expression:
Q ¼ SðKH  KC Þ
where KH and KC are the thermal conductivity integrals at the evaluated temperature
TH and TC, respectively. These two terms are defined as:
Z TH Z TC
KH ¼ ðkt dtÞ and KC ¼ ðkt dTÞ
0 0

and S is defined as a conduction shape factor as


Z x2
1=S ¼ ðdx=AðxÞÞ
x1

where kt is the material thermal conductivity at some temperature within the


temperature field expressed by the integral; dx is the change in length along the
axis of heat transfer; and A is the cross-sectional area of heat transfer.
The quantities KH and KC represent the integrated thermal conductivity for
any solid material that has had its thermal conductivity variation with temperature

© 2005 by Marcel Dekker


5367-4 Flynn Ch02 R2 090804

56 Chapter 2

determined. The values of KH and KC are also useful when their difference is divided
by the difference of TH and TC and becomes the apparent thermal conductivity of
the material, expressed as

KM ¼ ðKH  KC Þ=ðTH  TC Þ
Figures 2.9 and 2.10 are graphical depictions of thermal conductivity integrals
for some metals and insulators, respectively.
The conduction shape factor S is derived from the Fourier equation and simply
places all the physical parameters of the shape of the material into a single integral.
This allows for various geometric shapes to be evaluated and conduction shape fac-
tors to be determined based on the desired physical dimension of interest, such as
wall thickness or sphere diameters. Table 2.5 gives some representative conduction
shape factors for various geometries of interest. These geometries are pictured in
Fig. 2.11.
With this simple expression it is possible to get a good engineering estimate
on the heat balance around various cryogenic components under steady-state
conditions

dT
Q ¼ kA ð2:47Þ
dx
where Q is the heat transferred, k the material thermal conductivity (normally a
function of temperature), A the area of the material through which the heat is trans-
ferred and dT=dx the change in temperature observed in the material in the x direc-
tion due to the heat transfer.
If Q, k, and A are constant, the relation becomes
T2  T1
Q ¼ kA ð2:48Þ
Dx

This relation is useful when determining the heat transfer through a layer of insula-
tion or a series of insulation materials. If the conduction is through insulation
wrapped around a pipe, the relationship is modified to

dT T1  T2
Q ¼ kA ¼ 2 p Lk ð2:49Þ
dr lnðr2 =r1 Þ

where L is the length of the pipe, r1, r2 the inside and outside radii of the insulation
and T1, T2 the corresponding temperatures at these locations.
For nonsteady-state conductive heat transfer, a heat balance may be made
around a volume element of material of dimensions dx, dy, and dz. In the x direc-
tion, the heat transferred into the volume element is given by

@T
dQx ¼ k dy dz ð2:50Þ
@x
The heat leaving the volume element is

@dQx
dQ0x ¼ dQx þ dx ð2:51Þ
@x
If k is constant, then

© 2005 by Marcel Dekker


5367-4 Flynn Ch02 R2 090804

Basic Principles 57

Figure 2.9 Thermal conductivity integrals of metals. Data provided from S. Courts,
Lakeshore Cryotronics.

@2T
dQ0x  dQx ¼ k dy dz dx ð2:52Þ
@x2
Similar expressions are valid for the other two directions. If heat is stored in the
volume element, then

© 2005 by Marcel Dekker


5367-4 Flynn Ch02 R2 090804

58 Chapter 2

Figure 2.10 Thermal conductivity integrals of insulating materials. Data provided from
S. Courts, Lakeshore Cryotronics.

dQs ¼ dQx  dQ0x þ dQy þ dQz  dQ0z ð2:53Þ


but
@T
dQs ¼ CP r dx dy dz ð2:54Þ
@y

© 2005 by Marcel Dekker


5367-4 Flynn Ch02 R2 090804

Basic Principles 59

Table 2.5 Example Conduction Shape Factorsa

Geometry Shape factor

[1] Plane wall or slab S ¼ A=Dx


2pL
[2] Hollow cylinder S¼
lnðD0 =D1 Þ
2pD0 Di
[3] Hollow sphere S¼
ðD0  Di Þ
2pL
[4] Isothermal cylinder placed vertically in a semi-infinite medium S¼
lnð4L=DÞ
2pL
[5] Isothermal cylinder buried in a semi-infinite medium S¼
cosh1 ð2H=DÞ
8pD
[6] Isothermal sphere buried in a semi-infinite medium S¼
1  ðD=4HÞ
2pL
[7] Cylinder centered in a large plate S¼
lnð4H=DÞ
a
See Fig. 2.11 for correlated sketch of referenced geometry.

so
       2 
@T @2T @2T @ T
¼a þ þ ð2:55Þ
@y @x2 @y2 @z2

where y is the time, CP and r are the heat capacity and density of the material,

Figure 2.11 Sketch of geometries referred to in Table 2.5.

© 2005 by Marcel Dekker


5367-4 Flynn Ch02 R2 090804

60 Chapter 2

respectively, and a is defined as k=rCP. At a steady state, @T=@y ¼ 0. The solution of


Eq. (2.55 ) involves either the separation of variables or the use of integral trans-
forms (usually Laplace transforms). For regular geometries, plots have been devel-
oped that permit rapid and reasonably reliable solutions even at low temperatures
for heating or cooling times.

3.2. Convective Heat Transfer


Thermal convection is the transfer of heat from a fluid to a colder surface by means
of fluid particle motion. There are two types of convective heat transfer. Forced con-
vection occurs when a fluid is forced or pumped past a surface; free convection or nat-
ural convection occurs when fluid motion is caused by density differences within the
fluid. In either case, the basic equation is

Q ¼ hc AðTs  Tb Þ ð2:56Þ

where hc is the individual heat transfer coefficient, Ts is the surface temperature, and
Tb is the bulk fluid temperature. The individual heat transfer coefficient is critical in
the solution of Eq. (2.56) and must generally be obtained with the aid of some
empirical relationship that has been found to correlate the convective heat transfer
reasonably well for a specific geometry or flow pattern. The empirical relationships
used to obtain the individual heat transfer coefficients at low temperatures are very
similar to the relationships used at ambient temperatures.
Tables 2.6–2.8 provide some useful heat transfer relationships for
low-temperature heat exchangers for a variety of operating conditions. The units
for the relationships in these tables are in consistent American engineering units.
Conversion factors to other systems of measurement are found in the Appendix.
If heat is transferred from one fluid to another separated by a pipe as in a heat
exchanger, then Eq. (2.56 ) can be modified to include the individual heat transfer
coefficients for both fluids in an overall heat transfer coefficient as

Q ¼ Uo Ao DToverall ð2:57Þ

where Uo is based on the outside surface area Ao. If the inside surface area is used,
the relationship can be modified to

Q ¼ Ui Ai DToverall ð2:58Þ

If the heat exchange is across a pipe, Ui can be obtained from

1
Ui ¼ ð2:59Þ
1=hi þ Drp Ai =kp Ap;lm þ Ai =ho Ao

where hi is the inside heat transfer coefficient, Drp is the thickness of the pipe, Ai is
the inside surface area, kp is the thermal conductivity of the pipe, Ap,lm is the log
mean area of the pipe given by (Ao  Ai)[ln (Ao=Ai)], Ao is the outside surface area,
and ho is the outside heat transfer coefficient. A similar expression can be written for
Uo. For a flat wall, Eq. (2.59 ) reduces to

1
Ui ¼ Uo ¼ ð2:60Þ
1=hi þ Dx=k þ 1=ho

© 2005 by Marcel Dekker


Basic Principles
Table 2.6 Empirical Heat Transfer and Corresponding Pressure Drop Relationships for Concentric Tube and Extended Surface Heat Exchangers

Empirical pressure
Type Flow conditions Empirical heat transfer equation drop=length equation
" #
0:0668ðDe =LÞRe Pr k DP 32 G 2
Straight tubular pipe Flow inside, Re < 2100; h ¼ 3:658 þ 2=3 D
¼
1 þ 0:04ðDe =LÞRe Pr e DL Re gc De r
no phase change (gas)
"  2=3 #  0:14
De CP mf DP 0:158 G2

5367-4 Flynn Ch02 R2 090804


Straight tubular pipe Flow inside, h ¼ 2:439  106 ðRe2=3  125Þ 1 þ ¼ 0:2
L De mb DL Re gc De r
2100 < Re < 10,000; no
phase change (gas)
DP 0:092 G2
Straight tubular pipe Flow inside, Re > 10,000; h ¼ 0.023 CPG Re0.2 Pr2=3 ¼ 0:2
DL Re gc De r
no phase change (gas)
DP 0:092 G2 ½1 þ 3:5ðDe =Dh Þ
Helical tubular pipe Flow inside, Re > 10,000; h ¼ 0.023 CPG Re0.2 Pr2=3[1 þ 3.5(De=Dh)] ¼
no phase change (gas)
DL Re0:2 gc De r

DP 0:952 G2
Collins tubing Flow in annulus, h ¼ 0.118 CPG Re0.2 Pr2=3 ¼ 0:2 max
DL Re gc De r
400 < Re < 10,000;
no phase change (gas)

For tubular pipe: De ¼ inside pipe diameter; for noncircular tubes, De ¼ 4(AcL=Aw), where Ac ¼ inside free-flow cross-sectional area, L ¼ length of pipe, and Aw ¼ heat transfer
or wetted tube surface area; for annulus, De ¼ D2  D1, where D2 ¼ inside diameter of outer pipe and D1 ¼ outside diameter of inner pipe; Dh ¼ diameter of helix.
For Collins tubing: De ¼ 4(Va=Aw), where Va ¼ active flow volume in annulus, Aw ¼ wetted area in annulus, including fins.
Evaluate fluid properties for all relations at the mean film temperature, Tm ¼ 0.5(Tw þ Tb), where Tw ¼ tube wall temperature and Tb ¼ bulk temperature of fluid. Constants in
these equations are for use with 51 units.

61
© 2005 by Marcel Dekker
62
Table 2.7 Empirical Heat Transfer and Corresponding Pressure Drop Relationships for Flows Normal to Outside of Tubes on Shell Side of Coiled Tube
Heat Exchangers

5367-4 Flynn Ch02 R2 090804


Empirical heat transfer
Type Flow conditions equation Empirical pressure drop=tube equation
DP h i
Banks of staggered tubes Flow normal to outside of h ¼ 0.33 CPG Re0.4 Pr2=3 ¼ 0:5 þ 0:235ðXT  1Þ1:08 Re0:15 Gmax
2
=gc r
DN
tubes, 2000 < Re < 3.2  104;
no phase change (gas)
DP
Banks of in-line tubes Flow normal to outside of h ¼ 0.26 CPG Re0.4 Pr2=3 ¼ ½0:088 þ 0:16XL ðXT  1Þn Re0:15 Gmax
2
=gc r
DN
tubes, 2000 < Re < 3.2  104;
no phase change (gas)

Re ¼ DoGmax=m, where Do ¼ outside diameter of tubes; Gmax ¼ m_ =Amin, with Amin ¼ minimum open free-flow area between tubes; XT ¼ transverse pitch=tube outside diameter;
N ¼ total number of tubes in line of flow; XL ¼ longitudinal pitch=tube outside diameter; n ¼ 0.43 þ (1.13=XL). Fluid properties for all relations are evaluated at the mean film
temperature, Tm ¼ 0.5(Tw þ Tb), where Tw ¼ tube wall temperature and Tb ¼ bulk temperature of fluid.

Chapter 2
© 2005 by Marcel Dekker
5367-4 Flynn Ch02 R2 090804

Basic Principles 63

Table 2.8 Empirical Heat Transfer and Corresponding Pressure Drop Correlations for
Brazed Aluminum Plate-Fin Heat Exchangersa

Empirical heat Empirical pressure


Type and flow conditions transfer equation drop=length equation

DP ð0:0099 þ 40:8 Re1:033 ÞG 2


Straight fins, 500 < Re < 104; h ¼ 0:0291 Re0:24 ¼
DL gc rDe
Pr2=3 CP G
h ¼ 7.87 mm (0.310 in.);
w ¼ 0.15 mm (0.006 in.);
492 fins=m (12.5 fins=in.);
no phase change (gas)
DP ð0:0834 þ 23:6 Re0:062 ÞG 2
Wavy fins, 300 < Re < 104; h ¼ 0:085 Re0:265 ¼
DL gc rDe
Pr2=3 CP G
h ¼ 9.53 mm (0.375 in.);
w ¼ 0.20 mm (0.008 in.);
591 fins=m (15 fins=in.);
no phase change (gas)
DP 7:04G2 Re0:547
Herringbone fins, h ¼ 0:555ðRe þ 500Þ0:482 ¼
DL gc rDe
400 < Re < 104; Pr2=3 CP G
h ¼ 10.82 mm (0.426 in.);
w ¼ 0.15 mm (0.006 in.);
472 fins=m (12 fins=in.);
no phase change (gas)
a
To match present vendor specifications, dimensions are also given in customary English units.
De ¼ 4(AcL=Aw), where Ac ¼ inside free-flow cross-sectional area, L ¼ length of pipe, and Aw ¼ heat trans-
fer or wetted tube surface area. Evaluate fluid properties for all relations at mean film temperature,
Tm ¼ 0.5(Tw þ Tb), where Tw ¼ tube wall temperature and Tb ¼ bulk temperature of fluid.

Heat transfer from a condensing vapor to a cooler surface may be considered


as a special case of convective heat transfer. Heat transfer calculations with conden-
sing vapors are generally made using a relation similar to Eq. (2.56 ), where Tb is the
temperature of the condensing vapor and Ts is the temperature of the surface upon
which condensation occurs. For filmwise condensation on the outside of horizontal
tubes, the convective heat transfer co-efficient can be obtained from
 1=4
k3 r2 ghfg
hc ¼ 0:725 ð2:61Þ
mf Do NDT
where hfg is the latent heat of vaporization, N is the number of rows of tubes in the
vertical plane, Do is the outside diameter of the tubes, and DT is the temperature
difference between the saturated vapor and the outside tube surface. For filmwise
condensation on the outside of vertical tubes, the convective heat transfer coefficient
relationship is modified to
 1=4
k3 r2 ghfg
hc ¼ 0:942 ð2:62Þ
mf LDT
Heat transfer from a surface to a boiling liquid can also be considered a special
case of convective heat transfer. There are commonly seen two types of boiling: nucle-
ate and film boiling. Within the nucleate boiling category, there are two distinct clas-
sifications: natural convection (also known as pool boiling) or forced convection

© 2005 by Marcel Dekker


5367-4 Flynn Ch02 R2 090804

64 Chapter 2

Figure 2.12 Pooling boiling depictions: (a) saturated conditions generate vapor and raises
the bulk fluid temperature while (b) subcooled conditions shown allows for no vapor genera-
tion but does raise in the bulk fluid temperature, as well.

boiling. There is a further subdivision of two primary boiling types (nucleate and film)
based upon whether the boiling fluid is under saturated or sub-cooled conditions. The
nucleate boiling conditions are depicted in Fig. 2.12 and 2.13, while 2.14 illustrates
the filming boiling process. Most experimental work at low temperatures has been
conducted on pool boiling. Such studies have shown that the boiling action at these
temperatures is also largely dependent on the temperature difference between the sub-
merged surface and the boiling liquid. Unfortunately, the available information on
boiling liquids at low temperatures is limited and also shows wide scatter in the
experimental data. Figure 2.15 shows typical experimental nucleate and film pool
boiling data for nitrogen at 1atm compared with predictive correlations. Figures
2.16–2.18 give similar data for oxygen, hydrogen, and helium. See ‘‘Suggested Read-
ing’’ at the end of the book for the original data sources called out on Figs. 2.15–2.18.

3.3. Radiative Heat Transfer


Thermal radiation involves the transfer of heat from one body to another at a lower
temperature by electromagnetic waves passing through the intervening medium.
Radiant energy striking a material may be partly absorbed, transmitted, or reflected.

Figure 2.13 Forced convection boiling: (a) subcooled conditions again allow for no vapor
generation while (b) saturated or bulk boiling generates vapor and any number of different
vapor-liquid flow conditions.

© 2005 by Marcel Dekker


5367-4 Flynn Ch02 R2 090804

Basic Principles 65

Figure 2.14 Film boiling occurs when a vapor covers the super-heated bottom surface: (a) if
conditions are subcooled, then no vapor will be generated, and (b) at saturated conditions
large amounts of vapor can be formed.

For some materials, for example glass or plastic, this can be written as

aþtþr¼1 ð2:63Þ

where a is the absorptivity, t is the transmissivity, and r is the reflectivity. For an


opaque material, a þ r ¼ 1, and for a blackbody, a ¼ 1. The emissive power E of a
unit surface is the amount of heat radiated by the surface per unit of time. The

Figure 2.15 Experimental and predictive nucleate and film boiling of nitrogen at 1 atm.

© 2005 by Marcel Dekker


5367-4 Flynn Ch02 R2 090804

66 Chapter 2

Figure 2.16 Experimental and predictive nucleate and film boiling of oxygen at 1 atm.

emissivity E of a surface is defined by


E ¼ E=Eb ð2:64Þ

where Eb is the emissive power of a blackbody at the same temperature. The emissive
power for a blackbody is given by
Eb ¼ sT 4 ð2:65Þ

Figure 2.17 Experimental and predictive nucleate and film boiling of hydrogen at 1 atm.

© 2005 by Marcel Dekker


5367-4 Flynn Ch02 R2 090804

Basic Principles 67

Figure 2.18 Experimental and predictive nucleate and film boiling of helium at 1 atm.

where s is the Stefan–Boltzmann constant. For a real surface, the relation is


modified to

E ¼ EsT 4 ð2:66Þ

When two bodies within visual range of each other are separated by a medium that
does not absorb radiation, an energy exchange occurs as a result of reciprocal pro-
cesses of emission and absorption. The net rate of heat transfer from one surface at
T1 to another surface at T2 can be calculated from

Q ¼ sA1 FA FE ðT14  T24 Þ ð2:67Þ

where A1 is the surface area of one of the bodies, FA is a shape and orientation factor
for the two bodies relative to area A1, and FE is the emission and absorptance factor
for the two bodies. If the surface of one body is small or enclosed by the surface of
the other body, FA ¼ 1; otherwise, FA must be evaluated. For a surface A1 radiating
to a large area or space, FE ¼ E1. For two large parallel plates with a small space
between them,

1
FE ¼ ð2:68Þ
1=E1 þ 1=E2  1

For a sphere of radius r1 inside another sphere of radius r2,

1
FE ¼ ð2:69Þ
1=E1 þ ðr1 =r2 Þ2 ½ð1=E2 Þ  1

© 2005 by Marcel Dekker


5367-4 Flynn Ch02 R2 090804

68 Chapter 2

For a long cylinder of radius r1 inside another cylinder of radius r2,


1
FE ¼ ð2:70Þ
1=E1 þ ðr1 =r2 Þ½ð1=E2 Þ  1

For the case where FA ¼ 1 and FE ¼ E1, we can define a radiative heat transfer coeffi-
cient as
 
s T14  T24
hr ¼ ð2:71Þ
T1 T1  T2

This can then be used to determine the heat transferred by radiation from the
relation

Q ¼ hr A1 ðT1  T2 Þ ð2:72Þ

4. MOMENTUM TRANSFER

Cryogenic fluids are often handled near their boiling point. Thus, flow is usually compli-
cated because vaporization occurs, resulting in two-phase flow. The transition from liquid
to vapor depends on the heat influx and pressure changes encountered in the system.
For single-phase flow, the most important point to recognize is the distinction
between laminar and turbulent flow. For low temperatures, the same concepts that
have been developed for ambient flow apply. For flows with Reynolds numbers less
than 2100, laminar flow is assumed to occur; and for flows with Reynolds numbers
greater than 3000, turbulent flow is assumed to occur. The range in between 2100
and 3000 is normally considered the transition region between laminar and turbulent
flow. Thus, the empirical relationships that have been used successfully in calculat-
ing, for example, the pressure drop in a pipe at ambient conditions are also valid
for systems using cryogenic fluids.
Tables 2.5–2.7 provide some useful relationships to determine the empirical heat
transfer coefficient and pressure drop per unit length of many common heat exchangers.
In cryogenic systems, two-phase flow cannot be avoided, particularly during
the cooldown of transfer lines. If a cooldown situation is encountered frequently,
complete system performance reliability requires design for this condition. It takes
only a slight localized heat influx or pressure drop to cause the evaporation of part
of the fluid. Vapor in such lines can significantly reduce the liquid transfer capacity.
This transition from liquid to vapor is usually an unequilibrated condition and
presents an undefined problem for design purposes.
The Martinelli–Nelson correlation has been found to be useful for estimating pres-
sure drop for a single-phase stream under steady-state flow conditions. However, this cor-
relation, developed for ambient conditions, has a tendency to underestimate the pressure
drop by as much as 10–30% because of the high tendency of cryogenic fluids to vaporize.
Another fluid problem encountered rather frequently is that of cavitation in
pumps. Cavitation begins at a pressure slightly below where vaporization just begins.
The difference in pressure between cavitation and normal vaporization is equal to the
net positive suction head minus the drop in head pressure in the pump. Normally, this
must be determined experimentally and is a characteristic of the pump and the fluid.
A critical flow can be achieved in the transfer of a cryogenic fluid. In such a
situation, the flow rate of the fluid cannot be altered by lowering the downstream

© 2005 by Marcel Dekker


5367-4 Flynn Ch02 R2 090804

Basic Principles 69

pressure as long as the supply pressure remains constant. Such a situation is due to
the increased down-stream vaporization, which reduces the cross-sectional area
available for liquid transfer.
Cooldown of cryogenic systems involves a period of surging flow and pressure
fluctuations with accompanying sound effects. The primary factors that affect surging
and maximum pressure include inlet pressure and temperature, pipe length and
diameter, pipe precooling, throttling, and liquid properties. Other factors such as
heat capacity, external heat leak, and pipe wall thickness have a minor effect on
surging, because the pipe temperature undergoes little change during such a surge.

5. HEAT LEAK AND PRESSURE DROP IN CRYOGENIC TRANSFER


LINES

Cryogenic fluid transfer presents a number of challenges to any designer but two
problem areas, in particular, have wide ranging impacts on the overall system design:
(1) heat leak in cryogenic transfer, and (2) pressure drop during cryogenic transfer of
fluids.
The transfer piping design should accommodate the desired heat transfer
requirements as well as the fluid transfer conditions necessary to accommodate the
overall system design. A comparison of the heat leak into a bare pipe vs. those with
various insulations techniques is given in Table 2.9 for liquid nitrogen. The effect of
improved insulation techniques, such as vacuum jacketed pipe (VJP), is obvious and
can result in reduced heat loads of several orders of magnitude over bare pipes. VJP
comes in rigid and flexible configurations to accommodate the designer’s need for
making bends, etc. The flexible configuration requires a more complex and slightly
less efficient insulation scheme and therefore do not offer the same performance as
the rigid versions but is still very good vs. the nonvacuum jacketed alternatives. Both
rigid and flexible transfer lines come in various sizes and a number of commercial
vendors are available to supply these types of units.
Table 2.10 gives estimated heat leak for liquid nitrogen, oxygen, and hydrogen
with various line sizes. An additional consideration of VJP is the required cooldown
heat removal necessary to allow the line to efficiently transfer cryogenic fluids. Note
that flow is not a parameter. This is the case because fully developed turbulent flow is
assumed. For laminar flow to exist, the pipe diameter would have to be extremely
large, resulting in excessive costs and unnecessarily high heat leak into the system.
At turbulent flow conditions, the controlling insulating parameter to heat leak into
the fluid is the vacuum jacket around the transfer pipe. Solid thermal conductivity
through the pipe walls, as well as the heat transfer coefficient to turbulent cryogenic

Table 2.9 Comparison of Insulation Techniques for Fluid


Transfer Piping Versus Bare or Non-insulated Piping Illustrates
Needs for Insulation (Courtesy Technifab Products, Inc.)

Pipe comparison Heat leak Btu=ft=hr

Bare copper pipe 200.0


Foam insulated pipe 20.0
Bare vacuum pipe 4.0
Vacuum jacketed pipe from TPI 0.47

© 2005 by Marcel Dekker


5367-4 Flynn Ch02 R2 090804

70 Chapter 2

Table 2.10 Estimate Heat Leak for Vacuum Jacketed Piping, Both Rigid and Flexible.
(Courtesy of PHPK Technologies, Inc.)

Vacuum jacketed piping heat leak in Btu=hr=ft and (Watts=ft)

LN2 LO2 LH2

Line Size Rigid Flex Rigid Flex Rigid Flex


3
4" OD  114" NPS 0.37 (0.11) 0.97 (0.28) 0.37 (0.11) 0.96 (0.28) 0.40 (0.12) 1.05 (0. 31)
3
4" NPS  2" NPS 0.43 (0.13) 1.21 (0.25) 0.42 (0.12) 1.19 (0.35) 0.46 (0.13) 1.29 (0.38)
1" NPS  212" NPS 0.47 (0.14) 1.43 (0.42) 0.47 (0.14) 1.41 (0.41) 0.51 (0.15) 1.54 (0.45)
112" NPS  3" NPS 0.58 (0.17) 1.74 (0.51) 0.57 (0.17) 1.71 (0.50) 0.63 (0.18) 1.89 (0.55)
2" NPS  4" NPS 0.79 (0.23) 2.37 (0.70) 0.65 (0.19) 1.95 (0.57) 0.85 (0.25) 2.56 (0.75)
3" NPS  5" NPS 0.98 (0.29) 2.95 (0.86) 0.84 (0.25) 2.52 (0.74) 1.08 (0.32) 3.24 (0.95)
4" NPS  6" NPS 1.28 (0.38) 3.85 (1.13) 1.01 (0.30) 3.03 (0.89) 1.40 (0.41) 4.22 (1.24)
6" NPS  8" NPS 1.65 (0.48) 4.97 (.146) 1.36 (0.40) 4.10 (1.20) 1.83 (0.54) 5.50 (1.61)

flow are both negligible to that of the radiation heat load through the vacuum space
per length of pipe. This radiation heat load is a first order effect relating to surface
areas, the larger the surface area, the larger the heat load. The internal and external
surface areas of pipes are driven by their respective diameters.
Table 2.10 also shows very little increase in heat leak in changing the fluid
transferred. All three fluids have heat leak conditions within about 10% of each other
even though the temperature changed from 90 to 20 K. From simple radiation heat
transfer calculations alone, one might expect a heat leak increase by a factor of over
400 (¼(90=20)4). However, the actual case is that the hydrogen temperatures do a
much better job of cryopumping than do oxygen or nitrogen cooled surfaces. Thus
at 90 K, there is still some small residual gas conduction, which is not present at
20 K. Therefore, the apparent thermal conductivities are not that much different.
Table 2.11 illustrates using vacuum jacket valves the cooldown loss in Btu for
various sizes at two temperatures, liquid nitrogen and liquid hydrogen, and then
compares the cooldown loss to the steady-state heat leaks of these units. The heat
leak into vacuum jacketed (VJ) valves is expected to be higher due to the (1) increase
of the penetrations into the VJ around the valve body, (2) the required increased

Table 2.11 Estimated Performance Data for Different Size Vacuum Jacketed Valves.
(Courtesy of PHPK Technologies, Inc.

Cool-down and heat leak performance information (jacketed)

Cool-down loss Btu Steady state heat leak Btu=hr (W)

Valve Size 80 K 20 K 80 K 20 K
1
2" 81 86 10.5 (3.1) 11.6 (3.4)
3
4" 81 86 10.5 (3.1) 11.6 (3.4)
1" 99 105 10.5 (3.1) 11.6 (3.4)
112" 189 200 13.7 (4.0) 15.0 (4.4)
2" 324 342 18.1 (5.3) 20.0 (5s.9)
3" 990 1045 39.7 (11.6) 44.1 (12.9)
4" 1260 1330 47.9 (14.0) 53.1 (15.6)

© 2005 by Marcel Dekker


5367-4 Flynn Ch02 R2 090804

Basic Principles 71

support for the internal valve components, (3) heat from the power source to drive
the valve, (4) instrumentation required to operate and monitor the valve, etc. Addi-
tional heat leak considerations should be included for the mechanical connections
between VJP segments and the components included in the fluid systems. Each indi-
vidual connection can be roughly estimated to be about 1.24 times a one-foot seg-
ment of the same size line.
The second design problem is pressure drop in cryogenic transfer lines. Two
figures (Figs. 2.19 and 2.20) are provided to show the expected pressure drop per foot
of pipe vs. flow in gallons per minute for nitrogen and hydrogen. Inside pipe dia-
meter is the parameter for the data lines. Additionally, Fig. 2.21 is provided and
illustrates the pressure drop differences of rigid pipe vs. flexible pipe for flowing
liquid nitrogen. The increase in pressure drop can be directly attributed to the inter-
nal design use of welded bellows to form the interior pipe walls. This is true in all, not
just cryogenic, fluid transfer involving flexible lines.
The values shown in Table 2.10 and the graphs of Figs. 2.19 and 2.20 can be
expected from currently available, commercially fabricated transfer lines using
current vacuum-jacketing technology. These values have been confirmed by the two
largest US manufacturers of this equipment. At their request, no credit is given, as
they have no control over how the data will be used and accordingly wish to assume
no responsibility for potential misuse by others.

Figure 2.19 Pressure drop of liquid nitrogen in Schedule 5 rigid pipe.

© 2005 by Marcel Dekker


5367-4 Flynn Ch02 R2 090804

72 Chapter 2

Figure 2.20 Pressure drop of liquid hydrogen in Schedule 5 rigid pipe.

Example 2.4. Given the following data, calculate the heat leak and pressure drop
for liquid hydrogen in a commercially available vacuum-insulated transfer line:
Liquid hydrogen transfer rate: 250 gpm
Maximum allowable heat leak: 800 Btu=hr
Acceptable pressure drop: 20–30 psi
Desirable pipe size: 2 in. i.d.
Heat leak. Table 2.10 shows that a 2 in. pipe will have a heat leak of 0.85 Btu=hr per
foot of pipe in which liquid hydrogen is flowing. The various fittings (spacers, valves,
bayonet connectors, etc.) will each amount to 1.24 equivalent feet of pipe.
Therefore, if 800 Btu=hr can be accommodated, the maximum run of equiva-
lent pipe is 800=0.85 ¼ 941 ft of pipe. If there were 24 components (a large number),
their equivalent length would be 24  1.24 ¼ 41 ft. Thus, the pipe run could comprise
900 ft of pipe and 24 components.
Pressure drop. Figure 2.20 for hydrogen flow shows a pressure drop of 0.034 psi per
foot of 2 in. pipe.
If the tolerated pressure drop is 20 psi, then the maximum length of pipe is
20=0.034 ¼ 588 ft of pipe.
If the acceptable pressure drop is 30 psi, then the maximum length of pipe is
30=0.034 ¼ 882 ft of pipe.

© 2005 by Marcel Dekker


5367-4 Flynn Ch02 R2 090804

Basic Principles 73

Figure 2.21 Pressure drop estimates for rigid and flexible transfer lines with liquid nitrogen.
(Courtesy of PHPK Technologies, Inc.).

Since heat transfer and pressure drop follow Rayleigh’s equivalence, a safe
assumption is that the pressure drop caused by a component is about 1.24 times that
of an equivalent length of pipe. Let us pick a pipe run of 500 ft of actual pipe with 24
components in it. The 24 components represent 24  1.24 ¼ 30 equivalent feet of
pipe. The heat leak for this case is then
ð530 ft of equivalent pipeÞ  ½0:85Btu=hr ftÞ ¼ 450:5 Btu total heat leak
The pressure drop for this case is
ð530 ft equivalent pipeÞ  ð0:034 psi=ftÞ ¼ 18 psi
Summary. For this sample case, 500 ft of 2 in. I.D. pipe carrying liquid hydrogen with
24 fittings will have a heat leak of 450 Btu and a pressure drop of 18 psi, which is well
within our stated parameters. Other cases are readily calculated in a similar manner.

6. COOLDOWN

The amount of liquid required for cooldown can be estimated rapidly by equating
the sensible heat of a pipeline and its insulation to the heat of vaporization of the
fluid. This, of course, does not consider heat leaking into the system, frictional losses,

© 2005 by Marcel Dekker


5367-4 Flynn Ch02 R2 090804

74 Chapter 2

or changing temperature of the gas discharge from the system. Should any one of
these items contribute a sizable heat load to the system, then the amount of liquid
required for cooldown would obviously be low. In the final analysis, economics will
normally be the most important factor in the selection of a piping system. Such
economics will depend largely on the duration of the transfer and the time
between transfers. These times must be evaluated rather closely if the economic
evaluation is to be valid.
In Chapter 4, we will see that the heat capacity of all crystalline materials goes
to zero as the absolute temperature goes to zero. Accordingly, the amount of cryo-
genic fluid needed for cooldown also goes to zero as the initial temperature of the
solid goes to zero. In other words, prechilling a liquid hydrogen system with liquid
nitrogen will save a lot of liquid hydrogen.
Figures 2.22–2.25 give cooldown values for nitrogen, oxygen, hydrogen, and
helium. They are based on the heat capacity of aluminum (Al); stainless steel (SS),
typically a 304 series; and copper. They all show that the amount of coolant per mass
of metal decreases rapidly as the initial cooldown temperature decreases. For each
metal, there is a pair of MAX and MIM curves. The MAX curve is the maximum
amount of coolant required if only the heat of vaporization of the liquid is available
to cool the metal. This is the worst case. The MIN curve is the minimum amount of
coolant required if both the heat of vaporization of the respective liquid and all of the
sensible heat capacity of the liquid turned gas are available to cool the metal. This is
the best case and will occur only if all of the boil-off gas is somehow in perfect heat
exchange with the metal being cooled, such that the boil-off gas vents at the same
temperature as the warm metal. The vent gases must move slowly past the metal
in order to achieve this ‘‘best case’’ cooling. In other words, all of the possible refrig-
eration is captured; none is wasted. Only by very careful and expensive design, such
as in the cooling of some space satellites, is the minimum ever achieved.

Figure 2.22 Cooldown to liquid nitrogen temperatures. Mass of liquid nitrogen required per
mass of metal vs. starting temperature.

© 2005 by Marcel Dekker


5367-4 Flynn Ch02 R2 090804

Basic Principles 75

Figure 2.23 Cooldown to liquid oxygen temperatures. Mass of liquid oxygen required per
mass of metal vs. starting temperature.

It is instructive to see just how much precooling a liquid hydrogen system with
liquid nitrogen can help. Referring to Fig. 2.2 for hydrogen, we see that starting at
540 R (80 F), it will take (MAX case) about 0.2 1b of liquid hydrogen to cool 1 lb of
stainless steel to 36 R (20 K). Now suppose that the stainless steel system is pre-
cooled with liquid nitrogen to 139 R (77 K). The same MAX curve now predicts that
only 0.01 1b of liquid hydrogen will be needed per pound of steel to cool it the rest of
the way to 36 R (77 K). The liquid hydrogen use has been cut by a factor of 20. The
price we pay (see Fig. 2.19) is that 0.4 1b of liquid nitrogen is used per pound of metal

Figure 2.24 Cooldown to liquid hydrogen temperatures. Mass of liquid hydrogen required
per mass of metal vs. starting temperature.

© 2005 by Marcel Dekker


5367-4 Flynn Ch02 R2 090804

76 Chapter 2

Figure 2.25 Cooldown to liquid helium temperatures. Mass of liquid helium required per
mass of metal vs. starting temperature.

for precooling from 540 R (300 K) to 139 R (77 K). Such a trade is very economical
indeed.

7. SUMMARY

As one would expect, the basic principles of thermodynamics, heat transfer, and
momentum transfer apply equally as well to cryogenic engineering as they do to
all other fields of engineering. Appropriate use must be made of the sometimes unu-
sual low-temperature properties of fluids and materials (see Chapter 3 and 4).
Simple engineering correlations, usually based on the behavior of a large
number of near-room-temperature fluids, must be used with caution. Pressure drop
correlations, for instance, are usually based on the flow of single-phase streams
under steady-state conditions. Thus they may underestimate the pressure drop in
cryogenic systems by 10–30% due to the high vaporization tendency of cryogenic
fluids.
Finally, the usual modeling assumptions for normal engineering work may not
apply. Cryogenic fluids are nearly always at their boiling point, are seldom adiabatic,
and are often composed of two phases. Radiative heat transfer can seldom be
neglected.
This is not to say that cryogenic systems cannot be rigorously modeled, for
they can be, if one takes into account all of the peculiar properties of cryogenic
materials and properties. Rather, correlations must be used with caution, and a
return to basics is usually the safest approach.

© 2005 by Marcel Dekker


5367-4 Flynn Ch03 R2 090804

3
Cryogenic Fluids

1. PVT BEHAVIOR OF A PURE SUBSTANCE

The relationship of specific or molar volume to temperature and pressure for a pure
substance in an equilibrium state can be represented by a surface in three
dimensions, as shown in Fig. 3.1. The shaded surfaces marked S, L, and G in
Fig. 3.1 represent, respectively, the solid, liquid, and gas regions of the diagram.
The unshaded surfaces are the regions of co-existence of two phases in equilibrium,
and there are three such regions: solid–gas (S–G), solid–liquid (S–L), and liquid–gas
(L–G).
Heavy lines separate the various regions and form boundaries of the surfaces
representing the individual phases. Because of the difficulty of visualizing and draw-
ing these surfaces, it is customary to depict the data on projections such as the PT
and PV planes. Figure 3.2 is an example of a PT plane for water; other materials
have similar diagrams.
The heavy line passing through points A and B marks the intersections of the
two-phase regions and is the three-phase line, along which solid, liquid, and gas
phases exist in three-phase equilibrium. According to the phase rule, such three-
phase, single-component systems have zero degrees of freedom; they exist for a given
pure substance at only one temperature and one pressure. For this reason, the pro-
jection of this line on the PT plane of Fig. 3.2 (shown to the left of the main diagram)
is a point. This is known as a triple point.
The triple point is merely the point of intersection of the sublimation and
vaporization curves. It must be understood that only on a P–T diagram is the triple
point represented by a point. (On a P–V diagram it is a line.)
The phase rule also requires that systems made up of two phases in equilibrium
have just one degree of freedom, and therefore the two-phase regions must project as
lines on the PT plane, forming a P–T diagram of three lines—fusion (or melting),
sublimation, and vaporization—which meet at the triple point.
The P–T diagram is divided into areas by three lines: the solid–vapor, solid–
liquid, and liquid–vapor lines. The triple point is a region in which liquid, solid,
and vapor can exist in equilibrium simultaneously.
The P–T projection of Fig. 3.2 is shown to a larger scale in Fig. 3.3.
Upon projecting the surface on the PT plane, the whole solid–vapor region
projects into the sublimation curve, the whole liquid–vapor region projects into
the vaporization curve, the whole solid–liquid region projects into the fusion curve,
and finally the ‘‘triple point line’’ projects into the triple point.

77

© 2005 by Marcel Dekker


5367-4 Flynn Ch03 R2 090804

78 Chapter 3

Figure 3.1 P–V–T surface of a pure substance.

The solid lines of Fig. 3.3 clearly represent phase boundaries. The fusion curve
(line 2–3) normally has a positive slope, but for a few substances (water is the best
known) it has a negative slope. This line is believed to continue upward indefinitely.
The two curves 1–2 and 2–C represent the vapor pressures of solid and liquid,
respectively.
The terminal point C is the critical point, which represents the highest pressure
and highest temperature at which liquid and gas can coexist in equilibrium. The
termination of the vapor pressure curve at point C means that at higher temperatures
and pressures no clear distinction can be drawn between what is called liquid and
what is called gas. At the critical point, there is no distinction between the liquid
and vapor phases.

Figure 3.2 PT projection of the P–V–T surface of a pure fluid.

© 2005 by Marcel Dekker


5367-4 Flynn Ch03 R2 090804

Cryogenic Fluids 79

Figure 3.3 P–T plot for a pure substance.

Homogeneous fluids are normally divided into two classes, liquids and gases.
However, the distinction cannot always be sharply drawn, because the two phases
become indistinguishable at the critical point.
Thus, there is a region extending indefinitely upward from and indefinitely to
the right of the critical point that is called simply the fluid region. The fluid region,
existing at higher temperatures and pressures, is marked off by dashed lines that do
not represent phase changes but depend on arbitrary definitions of what constitutes
liquid and gas phases.
The region designated as liquid in Fig. 3.3 lies above the vaporization curve.
Thus, a liquid can always be vaporized by a sufficient reduction in pressure at
constant temperature. The gas region of Fig. 3.3 lies to the right of the sublimation
and vaporization curves. Thus, a gas can always be condensed by a sufficient
reduction in temperature at constant pressure. Because the fluid region fits neither
of these definitions, it can be called neither a gas nor a liquid.
The gas region is sometimes considered to be divided in two parts, as shown by
the dash-dot vertical line of Fig. 3.3. A gas to the left of this line, which can be
condensed either by compression at constant temperature or by cooling at constant
pressure, is frequently called a vapor. A vapor is a gas existing at temperatures below
Tc, and it may therefore be condensed either by reduction of temperature at constant
P or by increase of pressure at constant T.
The P–T projection of Fig. 3.3 provides no information about the volumes of the
systems represented. This is, however, given explicitly by the projection of Fig. 3.4,
where all surfaces of the three-dimensional diagram show up as areas on the PV plane.
The liquid and gas regions of the PV projection are shown in more detail by
Fig. 3.5.
Since the PV plane is perpendicular to the PT plane in the P,V,T, coordinate
system, the solid–vapor, liquid–vapor, and solid–liquid areas of the PV plane appear
as lines on the PT plane or phase diagram.
The primary curves of Fig. 3.5 give the pressure–volume relations for saturated
liquid (A–C) and for saturated vapor (C–B). The area lying below the curve ACB

© 2005 by Marcel Dekker


5367-4 Flynn Ch03 R2 090804

80 Chapter 3

Figure 3.4 Projection of the P-V-T surface of a pure fluid on PV and PT planes.

represents the two-phase region where saturated liquid and saturated vapor coexist at
equilibrium. Point C is the critical point, for which the coordinates are Pc and Vc.
A number of constant-temperature lines or isotherms are included on Fig. 3.5.
The critical isotherm at temperature Tc passes through the critical point. Isotherms

Figure 3.5 P–V plot for a pure substance.

© 2005 by Marcel Dekker


5367-4 Flynn Ch03 R2 090804

Cryogenic Fluids 81

for higher temperatures T1 and T2 lie above the critical isotherm. Such isotherms do
not cross a phase boundary and are therefore smooth.
Lower temperature isotherms T3 and T4 lie below the critical isotherm and are
made up of three sections. The middle section traverses the two-phase region and is
horizontal, because equilibrium mixtures of liquid and vapor at a fixed temperature
have a fixed pressure, the vapor pressure, regardless of the proportions of liquid and
vapor present. Points along this horizontal line represent different proportions of
liquid and vapor, ranging from all liquid at the left end to all vapor at the right.
The locus of these end points is represented by the dome-shaped curve labeled
ACB, the left half of which (from A to C) represents saturated liquid, and the right
half (from C to B), saturated vapor.
The horizontal line segments become progressively shorter with increasing tem-
perature, until at the critical point the isotherm exhibits a horizontal inflection. As
this point is approached, the liquid and vapor phases become more and more nearly
alike, eventually becoming unable to sustain a meniscus between them and ultimately
becoming indistinguishable.
The sections of isotherms to the left of line AC traverse the liquid region and
are very steep, because liquids are not very compressible; that is, it takes large
changes in pressure to effect a small volume change. A liquid that is not saturated
(not at its boiling point, along line AC) is often called subcooled liquid or
compressed liquid.
The sections of isotherms to the right of line CB lie in the vapor region. Vapor
in this region is called superheated vapor to distinguish it from saturated vapor as
represented by the line CB.

2. TEMPERATURE–ENTHALPY AND TEMPERATURE–ENTROPY


DIAGRAMS OF PURE SUBSTANCES

Consider 1 lb of solid nitrogen encased in a cylinder fitted with a weighted piston


such that the pressure is always constant at P1.
Let the solid nitrogen be heated from a very low temperature (40 or 50 K). The
temperature of the solid nitrogen ice will rise as the nitrogen ice absorbs the energy
until the melting point (B) is reached. This process is represented on the T–H and
T–S diagrams of Figs. 3.6 and 3.7 by line AB. The energy added is represented on
the T–H diagram by the change in enthalpy (H) between R A and B and by the area
under the line AB on the T–S diagram, since Q ¼ T ds.
As more energy is added to the nitrogen ice, it will melt at constant pressure and
form a subcooled liquid at point C. Additional heat causes the liquid to increase in
temperature until it reaches the saturation temperature at point D. As vapor
is formed, the amount of liquid decreases until all liquid disappears and satura-
ted vapor fills the cylinder, corresponding to point F. At some intermediate point such
as E, the mixture of liquid and vapor is referred to as wet vapor having a quality X.
If the heating process described in the preceding paragraphs is repeated at var-
ious pressures, a series of constant-pressure lines similar to line P1 may be obtained.
The locus of all B points forms the saturated solid curve, while the locus of all D and
F points forms the saturated liquid and saturated vapor curves, respectively.
Figure 3.8 gives the generalized layout of a typical T–S chart for any fluid
except helium. The solid lines are phase boundaries. The region labeled ‘‘Liquid I’’
is the homogeneous liquid region. This region is bounded on the left by the

© 2005 by Marcel Dekker


5367-4 Flynn Ch03 R2 090804

82 Chapter 3

Figure 3.6 T–H diagram for a pure substance.

solid=liquid melting curve, which extends indefinitely, almost vertically. Liquid I is


bounded on the right by the saturated liquid=vapor line of the liquid–vapor II
region. The horizontal phase boundary at the bottom is the solid locus. The three
phase boundaries meet at the triple point, which is a single fixed temperature. Since
this projection shows specific volume information, the triple point shows as the line
that it really is, the triple point line.
Every fluid except helium has a characteristic T–S diagram showing the regions
where the three phases exist. For many fluids, the solid region is either not of interest
or not of enough interest to pay for measuring its boundaries. In this, the usual case,
only T–S information above the triple point will be displayed. Figure 3.9 shows the
complete T–S region for nitrogen, with all three phases. It is a bit unusual to find a
thermodynamic diagram with all three phases present, but that third phase and the
triple point line exist nonetheless.

Figure 3.7 T–S diagram for a pure substance.

© 2005 by Marcel Dekker


5367-4 Flynn Ch03 R2 090804

Cryogenic Fluids 83

Figure 3.8 T–S diagram phase boundaries near the triple point.

3. PROPERTIES AND USES OF CRYOGENIC FLUIDS

The cryogenic region of most interest is characterized principally by five fluids:


oxygen, nitrogen, neon, hydrogen, and helium. We do, in fact, speak of the ‘‘oxygen
range’’ or the ‘‘hydrogen range.’’ Table 3.1 gives the normal (0.987 bar, or 1 atm)
boiling temperature, the normal melting temperature, the critical temperature and
pressure, and the normal latent heat of vaporization for these five cryogenic fluids
and several other common cryogens (Tables 3.2–3.6).
The liquid temperature ranges of several cryogenic fluids are shown in Fig. 3.10.
The bottom point () is the triple point; the midpoint (þ) is the normal boiling point
(NBP), and the topmost point (G) is the critical point. Note that there is no conve-
nient fluid to provide refrigeration between the critical point of neon, 44.5 K, and the
freezing point of nitrogen, 63.5 K. The chart shows that, in principle, subatmo-
spheric pressure oxygen could be used. In reality, liquid oxygen is considered to
be much too hazardous for use in refrigeration. In addition, the subatmospheric
pressure invites undetected leaks into the dewar, a most dangerous situation indeed.
Figure 3.10 also shows that, in general, the liquid range increases as the NBP
increases. Oxygen and propylene ðC¼ 3 Þ, are anomalous in this respect. Their liquid
ranges are much greater than the liquid ranges of their neighbors, no doubt because
of their chemical bond hyperactivity. The freezing point of methane is tantalizingly
close to the NBP of oxygen, indicating that methane might be soluble in liquid oxy-
gen. This is indeed the case, as proved by McKinley and again recently by Flynn. In
addition, Flynn has shown that such mixtures are not overly sensitive to shock and
can be handled safely with about the same precautions as any other rocket fuel.

© 2005 by Marcel Dekker


84

5367-4 Flynn Ch03 R2 090804


Chapter 3
Figure 3.9 T–S diagram for nitrogen near the triple point.

© 2005 by Marcel Dekker


Cryogenic Fluids
Table 3.1 Selected Properties of Cryogenic Liquids at Their Normal Boiling Points

Property Oxygen Nitrogen Neon e-Hydrogena Helium-4 Air Fluorine Argon Methane

Normal boiling point (K) 90.18 77.347 27.09 20.268 4.224 78.9 85.24 87.28 111.7
Density (kg=m3) 1141 808.9 1204 70.78 124.96 874 1506.8 1403 425.0
Heat of vaporization (kJ=kg) 212.9 198.3 86.6 445.6 20.73 205.1 166.3 161.6 511.5
Specific heat (kJ=(kg K)) 1.70 2.04 1.84 9.78 4.56 1.97 1.536 1.14 3.45
Viscosity (kg=(m s)  106) 188.0 157.9 124.0 13.06 3.57 168 244.7 252.1 118.6
Thermal conductivity(mW=(m K)) 151.4 139.6 113 118.5 27.2 141 148.0 123.2 193.1
Dielectric constant 1.4837 1.434 1.188 1.226 1.0492 1.445 1.43 1.52 1.6758
Critical temperature (K) 154.576 126.20 44.4 32.976 5.201 133.3 144.0 150.7 190.7
Critical pressure (MPa) 5.04 3.399 2.71 1.293 0.227 3.90 5.57 4.87 4.63
Temperature at triple point (K) 54.35 63.148 24.56 13.803 – – 53.5 83.8 88.7
Pressure at triple point (MPa  103) 0.151 12.53 43.0 7.042 – – 0.22 68.6 10.1
a
For a discussion of forms of hydrogen, see Sec. 3.7.

85
© 2005 by Marcel Dekker
Table 3.2 Properties of Cryogenic Liquids at Their NBP, Critical and Triple Points

86
Property Helium Hydrogen Neon Nitrogen Air Fluorine Argon Oxygen Methane

Molecular weight 4.0026 2.01594 20.1790 28.0134 28.975 37.9968 39.9480 31.9988 16.0420
Normal boiling point, K 4.2 20.4 27.1 77.3 78.9 84.95 87.3 90.2 111.7
Normal boiling point, R 7.57 36.7 48.8 139.2 141.8 152.9 157.1 162.4 201.1
NBP Liquid density, kg=m3 124.9 70.9 1204 810.8 874.0 1502 1403 1134 425.0
NBP Liquid density, lbm=ft3 7.80 4.43 75.17 50.61 54.56 93.7 87.56 70.8 26.53
NBP Vapor density, kg=m3 16.845 1.3395 9.5924 4.624 4.487 5.629 5.7765 4.4668 1.8176
NBP Vapor density, lbm=ft3 1.0516 0.08362 0.59883 0.28866 0.28011 0.3513 0.36062 0.27885 0.11346
NBP Heat of vaporization, kJ=kg 20.7 446.3 85.71 198.8 205.1 157.4 161.6 213.1 510.33
NBP Heat of vaporization, Btu=lbm 8.92 191.9 36.849 85.32 88.2 67.8 69.5 91.63 219.2
Critical temperature, K 5.2 33.2 44.4 126.12 133.3 144.31 150.9 154.6 190.7
Critical temperature, R 9.36 59.74 79.9 227.0 240 259.8 271.2 278.3 343.3
Critical pressure, MPa 0.23 1.31 2.71 3.38 3.90 5.22 4.87 5.06 4.63
Critical pressure, Atm 2.26 12.98 26.84 33.5 38.7 51.47 48.3 50.1 45.8
Triple point, K 2.177(l) 13.9 24.56 63.2 60.6 53.5 83.8 54.4 88.7
Triple point, R 3.92(l) 25.1 44.2 113.7 109.1 96.37 150.8 98.0 159.7
Pressure at triple point (MPa)103 4.73(l) 7.2 43.0 12.8 7.035 0.25 68.6 0.15 10.1
Pressure at triple point, atm 0.0469(l) 0.0711 0.426 0.1268 0.0697 0.00249 0.679 0.00150 0.099
Gas density NTP 1 atm, 70 F
r lb=f3 0.010343 0.0052089 0.052148 0.072445 0.07493 0.09803 0.10355 0.082787 0.041552
r kg=m3 (g=L) 0.16569 0.083438 0.83532 1.1604 1.2000 1.5704 1.6555 1.3261 0.66559

Volume ratio

GAS NTP; 1 atm 70 F 754.2 848.5 1433.7 695.3 728.1 956.7 842.0 860.6 634.7
Liquid NBP

Ft3 GAS NTP; 1 atm 70 F 26.63 29.96 50.9 24.55 25.71 33.78 29.70 30.39 22.41

Chapter 3
Liter Liq. NBP

© 2005 by Marcel Dekker


5367-4 Flynn Ch03 R2 090804

Cryogenic Fluids 87

Table 3.3 Enthalpy of Phase Changes and Density of Five Common Cryogenic Fluids

Fluid He P-H2 N2 O2 C1
Molecular weight 4.0026 2.01594 28.0134 31.9988 16.0420

Triple point (l)


T, K 2,176.8 13.800 63.148 54.359 90.685
T,  F 455.75 434.83 346.00 361.82 296.44
P, atm 0.04976 0.06950 0.12356 0.001461 0.11543
P, psi 0.73125 1.0214 1.8159 0.021465 1.6964
rl, lb=f3 9.1239 4.8093 54.294 81.536 28.187
rl, kg=m3 (g=l) 146.15 77.038 869.70 1306.1 451.51
rv, lb=ft3 – 0.0079284 0.00421 0.0006715 0.015732
rv, kg=m3 (g=l) – 0.127 0.672 0.010756 0.252
rs, lb=ft3 – 5.43 (62.4) 84.82 30.41
rs, kg=m3 (g=l) – 86.5 (1000) 1358.7 487.1
DHmelt, BTU=lb – 25.03 (11.0) 5.98 27.04
DHmelt, J=gm – 58.23 (25.6) 13.908 62.9
DHsubl, BTU=lb – 218.14 103.6 111.5 260.1
DHsubl, J=gm – 507.39 241.0 259.3 605
DHvap, BTU=lb – 192.97 92.602 104.35 233.52
DHvap, J=gm – 448.85 215.39 242.72 543.16
Normal boiling point
T, K 4.2219 20.277 77.313 90.191 111.69
T,  F 452.07 423.17 320.51 297.33 258.63
rl, lb=f3 7.8004 4.4197 50.367 71.240 26.372
rl, kg=m3 (g=l) 124.95 70.797 806.79 1141.2 422.44
rv, lb=ft3 1.0516 0.08362 0.28866 0.27885 0.11347
rv, kg=m3 (g=l) 16,845 1.3395 4.6240 4.4668 1.8176
DHv BTU=lb 8.9081 191.51 85.496 91.597 219.40
DHv J=gm 20.720 445.44 198.86 213.05 510.33
Critical point
T, K 5.1953 32.938 126.19 154.58 190.55
T,  F 450.32 400.38 232.52 181.42 116.68
P, atm 2.2453 12.670 33.534 49.771 45.391
P, psia 32.996 186.20 492.81 731.43 667.06
r, lb=f3 4.3476 1.9577 19.547 27.228 10.154
r, kg=m3 (g=l) 69.641 31.360 313.11 436.14 162.65
NTP 1 atm, 70 F
r, lb=f3 0.010343 0.0052089 0.072445 0.082787 0.041552
r, kg=m3 (g=l) 0.16569 0.083438 1.1604 1.3261 0.66559
Volume ratio

GAS NTP; 1 atm 70 F 754.2 848.5 695.3 860.6 634.7
Liquid NBP

Ft3 GAS NTP; 1 atm 70 F 26.63 29.96 24.55 30.39 22.41
Liter Liq. NBP

Figure 3.11 shows the vapor pressure of several cryogenic liquids. Again, the
unusual liquid range of oxygen is shown.
Figure 3.12 gives the unusual correlation between the NBP and critical tem-
perature for several cryogenic fluids. In the absence of data, a rule of thumb is that
Tc is about 1.6 times TNBP or that the NBP is about 60% of the critical temperature.
Helium, of course, marches to its own drummer. It is surprising that the other quan-
tum fluids, hydrogen and neon, are so well behaved.

© 2005 by Marcel Dekker


88
Table 3.4 Triple Point, Normal Boiling Point and Critical Point Properties of Five Common Cryogenic Fluids

Density

T P Liquid Vapor Solid

MW K 
F 
C atm psi mm lb=ft3 g=L lb=ft3 g=L lb=ft3 g=L

Triple point (l He)


He 4.0026 2.1768 –455.75 –270.97 0.04976 0.73125 37.82 9.1239 146.15 – –
p-H2 2.01594 13.80 434.38 254.35 0.06950 1.0214 52.822 4.8093 77.038 0.0079284 0.127 5.43 86.5
N2 28.0134 63.148 346.00 210.00 0.12356 1.8159 93.909 54.294 869.70 0.00421 0.672 56.19 900
O2 31.9988 54.359 361.82 218.77 0.001461 0.021465 1.1101 81.536 1306.1 0.0006715 0.010756 84.82 1358.7
CH4 16.0420 90.685 296.44 182.47 0.11543 1.6964 87.729 28.187 451.51 0.015732 0.252 30.75 492

Density

T Liquid Vapor Heat of vaporization

MW K 
F 
C lb=ft3 g=L lb=ft3 g=L BTU=lb J=g

Normal boiling point


He 4.0026 4.2219 452.07 268.93 7.8004 124.95 1.0516 16.845 8.9081 20.720
p-H2 2.01594 20.277 423.17 252.87 4.4197 70.797 0.08362 1.3395 191.51 445.44
N2 28.0134 77.313 320.51 195.84 50.367 806.79 0.28866 4.6240 85.496 198.86

Chapter 3
O2 31.9988 90.191 297.33 182.76 71.240 1141.2 0.27885 4.4668 91.597 213.05
CH4 16.0420 111.69 258.63 161.46 26.372 422.44 0.11347 1.8176 219.40 510.33

© 2005 by Marcel Dekker


Cryogenic Fluids
T P Density

MW K 
F 
C atm psia lbs=ft3 g=l

Critical point
He 4.0026 5.1953 450.32 267.95 2.2453 32.996 4.3476 69.641
p-H2 2.01594 32.938 400.38 240.21 12.670 186.20 1.9577 31.360
N2 28.0134 126.19 232.52 146.96 33.534 492.81 19.547 313.11
O2 31.9988 154.58 181.42 118.57 49.771 731.43 27.228 436.14
CH4 16.0420 190.55 116.68 82.6 45.391 667.06 10.154 162.65

89
© 2005 by Marcel Dekker
5367-4 Flynn Ch03 R2 090804

90 Chapter 3

Table 3.5 Weight Volume Equivalents for Air, Argon, Nitrogen, Oxygen, Helium and
Hydrogen

Volume of liquid at Volume of gas at 70 F


normal boiling point and 14.696 PSIA

Weight lb Gallons L Cu ft Cu M L

Air 1.000 0.1371 0.5190 13.35 0.3779 377.9


7.294 1.000 3.785 97.33 2.756 2756.
1.927 0.2642 1.000 25.71 0.7281 728.1
7.493 1.027 3.889 100.0 2.832 2832.
2.646 0.3628 1.373 35.31 1.000 1000.
Argon 1.000 0.08600 0.3255 9.671 0.2739 273.9
11.63 1.000 3.785 112.5 3.184 3184.
3.072 0.2642 1.000 29.71 0.8412 841.2
10.34 0.8893 3.366 100.0 2.832 2832.
3.652 0.3141 1.189 35.31 1.000 1000.
Nitrogen 1.000 0.1482 0.5612 13.80 0.3908 390.8
6.745 1.000 3.785 93.11 2.637 2637.
1.782 0.2642 1.000 24.60 0.6965 696.5
7.245 1.074 4.066 100.0 2.832 2832.
2.559 0.3792 1.436 35.31 1.000 1000.
Oxygen 1.000 0.1050 0.3973 12.08 0.3420 342.0
9.527 1.000 3.785 115.0 3.258 3258.
2.517 0.2642 1.000 30.39 0.8606 860.6
8.281 0.8692 3.290 100.0 2.832 2832.
2.924 0.3070 1.162 35.31 1.000 1000.
Helium 1.000 0.9593 3.361 96.71 2.739 2739.
1.042 1.000 3.785 100.8 2.855 2855.
0.2754 0.2642 1.000 26.63 0.7542 754.2
1.034 0.9919 3.755 100.0 2.832 2832.
0.3652 0.3503 1.326 35.31 1.000 1000.
Hydrogena 1.000 1.693 6.409 192.0 5.436 5436.
0.5906 1.000 3.785 113.4 3.210 3210.
0.1560 0.2642 1.000 29.95 0.8481 848.1
0.5209 0.8821 3.339 100.0 2.832 2832.
0.1840 0.3115 1.179 35.31 1.000 1000.
a
Hydrogen data are for equilibrium hydrogen (99.79% para, 0.21% ortho).

Volume Equivalents

Density Volume ratio

1 atm, 70 F
Gas, 1 atm 70 F Ft3 Gas 1 atm, 70 F
3
MW lb=ft g=L Liquid NBP Liter liquid NBP

He 4.0026 0.010343 0.16569 754.2 26.63


p-H2 2.01594 0.0052089 0.083438 848.5 29.96
O2 28.0134 0.07244 1.1604 695.3 24.55
N 31.9988 0.082787 1.3261 860.6 30.39
CH4 16.0420 0.041552 0.66559 634.7 22.41

© 2005 by Marcel Dekker


5367-4 Flynn Ch03 R2 090804

Cryogenic Fluids 91

Table 3.6 Weight Volume Equivalents of Commercially Important Cryogenic Fluids


(courtesy of PHPK Technologies)

Weight of liquid or gas Volume of liquid or normal boiling point Volume of gas at 70 F

lb kg cu ft l quarts gallons cu ft cu m

Oxygen
1.000 0.454 0.0140 0.396 0.419 0.105 12.08 0.342
2.205 1.000 0.0308 0.873 0.923 0.231 26.63 0.754
71.5 32.43 1.000 28.32 29.92 7.481 863.5 24.45
2.524 1.145 0.0353 1.000 1.056 0.264 30.48 0.863
2.388 1.083 0.0334 0.946 1.000 0.250 28.84 0.817
9.560 4.336 0.134 3.786 4.000 1.000 115.5 3.271
8.28 3.756 0.116 3.285 3.472 0.868 100.0 2.832
2.92 1.324 0.0408 1.155 1.221 0.305 35.31 1.000
Air
1.000 0.454 0.0183 0.518 0.548 0.137 13.35 0.379
2.205 1.000 0.0404 1.144 1.209 0.302 29.44 0.834
54.57 24.77 1.000 28.32 29.92 7.481 728.5 20.63
1.927 0.875 0.0353 1.000 1.056 0.264 25.73 0.729
1.824 0.828 0.0334 0.946 1.000 0.250 24.35 0.690
7.294 3.311 0.134 3.786 4.000 1.000 97.37 2.758
7.493 3.402 0.137 3.880 4.099 1.025 100.0 2.832
2.646 1.201 0.0485 1.374 1.449 0.363 35.31 1.000
Argon
1.000 0.454 0.0114 0.323 0.341 0.0853 9.681 0.274
2.205 1.000 0.0252 0.714 0.755 0.189 21.35 0.605
87.4 39.6 1.000 28.32 29.92 7.481 846.1 23.96
3.085 1.399 0.0353 1.000 1.056 0.264 29.86 0.846
2.919 1.324 0.0334 0.946 1.000 0.250 28.26 0.800
11.71 5.310 0.134 3.786 4.000 1.000 113.4 3.212
10.33 4.686 0.118 3.342 3.533 0.883 100.0 2.832
3.648 1.655 0.0417 1.181 1.248 0.312 35.31 1.000
Nitrogen
1.000 0.454 0.0198 0.561 0.593 0.148 13.79 0.391
2.205 1.000 0.0437 1.239 1.310 0.327 30.41 0.861
50.4 22.86 1.000 28.32 29.92 7.481 695.2 19.69
1.78 0.807 0.0353 1.000 1.056 0.264 24.55 0.695
1.68 0.762 0.0334 0.946 1.000 0.250 23.17 0.656
6.75 3.062 0.134 3.786 4.000 1.000 93.10 2.637
7.25 3.289 0.144 4.078 4.310 1.077 100.0 2.832
2.56 1.161 0.0508 1.439 1.521 0.380 35.31 1.000
Helium
1.000 0.454 0.128 3.625 3.832 0.958 97.08 2.749
2.205 1.000 0.283 8.00 8.468 2.117 214.1 6.063
7.80 3.541 1.000 28.32 29.92 7.481 757.3 21.45
0.2753 0.125 0.0353 1.000 1.056 0.264 26.73 0.757
0.2605 0.118 0.0334 0.946 1.000 0.250 25.29 0.716
1.045 0.474 0.134 3.786 4.000 1.000 101.5 2.875
1.033 0.469 0.132 3.738 3.948 0.987 100.0 2.832
0.3637 0.165 0.0466 1.320 1.396 0.349 35.31 1.000

(Continued)

© 2005 by Marcel Dekker


5367-4 Flynn Ch03 R2 090804

92 Chapter 3

Table 3.6 (Continued )

Weight of liquid or gas Volume of liquid or normal boiling point Volume of gas at 70 F

lb kg cu ft l quarts gallons cu ft cu m

Hydrogen
1.000 0.454 0.226 6.400 6.764 1.691 191.6 5.426
2.205 1.000 0.498 14.10 14.90 3.726 422.4 11.96
4.43 2.011 1.000 28.32 29.92 7.481 848.7 24.04
0.156 0.0708 0.0353 1.000 1.056 0.264 29.89 0.846
0.148 0.0672 0.0334 0.946 1.000 0.250 28.35 0.803
0.594 0.270 0.134 3.786 4.000 1.000 113.8 3.223
0.522 0.237 0.118 3.342 3.532 0.883 100.0 2.832
0.184 0.0835 0.0416 1.178 1.244 0.311 35.31 1.000

It is also possible to estimate the Joule–Thomson inversion temperature, as


shown in Fig. 3.13. Ignore the quantum fluids for a moment, and we see that the
J–T inversion temperature is about eight times the NBP. Including helium, hydro-
gen, and neon, the approximation is about 9 or 10 times the NBP.
Some of the important characteristics of the most widely used cryogenic liquids
are discussed more specifically in the following sections. For a comprehensive listing
of fluid properties, see Johnson (1960).

3.1. Oxygen
Oxygen was the base used for chemical atomic weights, being assigned the atomic
weight 16.000, until 1961, when the International Union of Pure and Applied Chem-
istry (IUPAC) adopted carbon-12 as the new basis. Oxygen has eight isotopes. Natu-
rally occurring oxygen consists of three stable isotopes of atomic mass numbers 16,
17, and 18, having abundances in the proportion 10,000:4:20. Oxygen condenses into
a light blue liquid whose density, 1134.2 kg=m3 (70.8 lbm=ft3) at its boiling tempera-
ture, is slightly greater than that of water at room temperature. At 0.987 bar (1 atm)
pressure, liquid oxygen boils at 90.2 K (162.4 R) and freezes at 54.4 K (97.9 R).
In the liquid state, there is thought to be some weak transient association of
oxygen molecules forming O4, which is said to be responsible for the blue color of
both liquid and solid oxygen. Ozone (O3), a highly active allotropic from of oxygen,
is formed by the action of an electrical discharge or ultraviolet light on oxygen.
Ozone’s presence in the atmosphere (amounting to the equivalent of a layer 3 mm
thick at ordinary pressures and temperatures) is of vital importance in preventing
harmful ultraviolet rays of the sun from reaching the earth’s surface.
Table 3.7 lists the temperature, pressure, and specific volume of each of the
various phases of oxygen at several important points. Tables 3.8–3.10 give properties
of oxygen in metric and English units at three of these fixed points.
Figures 3.14 and 3.15 and T–S diagrams for oxygen. A Mollier chart (pressure
vs. heat content) is given in Fig. 3.16, and the Joule–Thomson inversion chart in Fig.
3.17. Table 3.11 presents data for the J–T curve.

3.1.1. Properties of Oxygen


a. Magnetism. Liquid oxygen is slightly magnetic, in contrast to the other
cryogenic fluids, which are nonmagnetic. Its paramagnetic susceptibility is 1.003 at

© 2005 by Marcel Dekker


Cryogenic Fluids
93
Figure 3.10 Liquid temperature range of different cryogens.

© 2005 by Marcel Dekker


5367-4 Flynn Ch03 R2 090804

94 Chapter 3

Figure 3.11 Vapor pressure curves. CP ¼ critical point; TP ¼ triple point; l ¼ lambda point
(see Sec. 3.8.1).

its normal boiling temperature, so a bath of liquid oxygen is perceptibly attracted by


a magnet. This characteristic has prompted the use of a magnetic field in a liquid
oxygen (LOX) dewar to separate the liquid and gaseous phases under zero gravity
conditions. Also, by measuring the magnetic susceptibility, small amounts of oxygen
can be detected in mixtures of other gases. This paramagnetic property has
been exploited in commercial instruments for detecting small amounts of oxygen
in other gases.

Figure 3.12 Ratio of the critical temperature to the NBP for various cryogens. For C terms,
see Fig. 3.11.

© 2005 by Marcel Dekker


5367-4 Flynn Ch03 R2 090804

Cryogenic Fluids 95

Figure 3.13 Ratio of the inversion temperature to the NBP.

b. Chemical Reactivity. Both gaseous and liquid oxygen are chemically


reactive, especially with hydrocarbon materials. Because of its chemical activity,
oxygen presents a serious safety problem. Several explosions have resulted from
the combination of oxygen and hydrocarbon lubricants. Although at 90 K most
chemical reaction rates are negligibly slow, a small amount of energy added under
the right conditions can cause an explosion in a system containing liquid oxygen
and a substance with which it combines chemically.

Table 3.7 Fixed Points for Oxygen

Critical point
T ¼ 154.576  0.010 K T ¼ 154.581  278.237 R
P ¼ 49.76  0.02 atm (731.4 psia; 5.043 MPa)
V ¼ 73.37  0.10 cm3=mol (0.03673 ft3=lbm; 436.2 kg=m3)
Normal boiling point
T ¼ 90.180  0.01 K T ¼ 90.188  162.324 R
P ¼ 1 atm (14.696 psia; 101.3 kPa)
V (liquid) ¼ 28.05  0.028 cm3=mol (0.01404 ft3=lbm; 1141 kg=m3)
V (vapor) ¼ 7150  7 cm3=mol (3.579 ft3=lbm; 4.48 kg=m3)
Normal melting point
T ¼ 54.362  0.001 K T ¼ 54.372  97.852 R
P ¼ 1 atm (14.696 psia; 101.3 kPa)
V (liquid) ¼ 24.49  0.024 cm3=mol (0.01226 ft3=lbm; 1307 kg=m3)
Triple point
T ¼ 54.351  0.001 K T ¼ 54.361  97.832 R
P ¼ (1.50  0.06)  103 atm (0.0220 psia; 151.7 Pa)
V (solid) ¼ 23.55  0.03 cm3=mol (0.01179 ft3=lbm; 1359 kg=m3)
V (liquid) ¼ 24.49  0.024 cm3=mol (0.01226 ft3=lbm; 1307 kg=m3)
V (vapor) ¼ 2.97  0.01  106 cm3=mol (1487 ft3=lbm; 10.8 g=m3)
Solid-solid transitions. There are two such transitions at atmospheric pressure:
at 43.8 K and at 23.9 K

Source: H. M. Roder and L. A. Weber (1972).

© 2005 by Marcel Dekker


5367-4 Flynn Ch03 R3 100504

96 Chapter 3

Table 3.8 Oxygen Physical Constants

Oxygen
Chemical Symbol: O2
CAS Registry Number: 7782-44-7
DOT Classification: Nonflammable gas
DOT Label: Oxidizer
Transport Canada Classification: 2.2 (5.1)
UN Number: UN 1072 (compressed gas); UN 1073 (refrigerated liquid)

US Units SI Units

Physical constants
Chemical formula O2 O2
Molecular weight 31.9988 31.9988
Density of the gas at 70 F (21.1 C) and 1 atm 0.08279 lb=ft3 1.326 kg=m3
Specific gravity of the gas at 70 F (21.1 C) 1.105 1.105
and 1 atm (air ¼ 1)
Specific volume of the gas at 70 F (21.1 C) 12.08 ft3=lb 0.7541 m3=kg
and 1 atm
Boiling point at 1 atm 297.33 F –182.96 C
Freezing point at 1 atm 361.80 F 218.78 C
Critical temperature 181.43 F 118.57 C
Critical pressure 731.4 psia 5043 kPa abs
Critical density 27.22 lb=ft3 436.1 kg=m3
Triple point 361.82 F at 218.79 C at
0.02147 psia 0.1480 kPa abs
Latent heat of vaporization at boiling point 91.7 Btu=lb 213 kJ=kg
Latent heat of fusion at 361.1 F (218.4 C) 5.959 Btu=lb 13.86 kJ=kg
melting point
Specific heat of the gas at 70 F (21.1 C) and
1 atm
CP 0.2197 Btu=(lb)( F) 0.9191 kJ=(kg)( C)
CV 0.1572 Btu=(lb)( F) 0.6578 kJ=(kg)( C)
Ratio of specific heats (CP=Cv) 1.40 1.40
Solubility in water, vol=vol at 32 F (0 C) 0.0491 0.0491
Density of the liquid at boiling point 9.52 lb=gal or 1141 kg=m3
71.23 lb=ft3
Density of the gas at boiling point 0.2799 lb=ft3 4.483 kg=m3
Gas=liquid ratio (gas at 70 F (21.1 C) and 860.5 860.5
1 atm, liquid at boiling point, vol=vol)

To ensure against such unwanted chemical reactions, systems using liquid


oxygen must be maintained scrupulously clean of any foreign matter. The phrase
‘‘LOX clean’’ in the space industry has come to be associated with a set of elaborate
cleaning and inspection specifications representing a near ultimate in large-scale
equipment cleanliness. (See Chapter 10.) Ordinary hydrocarbon lubricants are
dangerous to use in oxygen compressors and vacuum pumps exhausting oxygen.
Also valves, fittings, and lines used with oil-pumped gases should never be used with
oxygen. Serious explosions have resulted from the combination of oxygen with a
lubricant. In fact, combustible materials soaked in liquid oxygen are used as inexpen-
sive commercial explosives.

© 2005 by Marcel Dekker


Table 3.9 Fixed Point Properties of Oxygen (Metric Units)

Cryogenic Fluids
Triple point Normal boiling point Standard conditions

Property Solid Liquid Vapor Liquid Vapor Critical point STP(0 C) NTP(20 C)

Temperature (K) 54.351 54.351 54.351 90.180 90.180 154.576 273.15 293.15
Pressure (mmHg) 1.138 1.138 1.138 760 760 37,823 760 760
Density (mol=cm3)  103 42.46 40.83 0.000336 35.65 0.1399 13.63 0.04466 0.04160
Specific volume (cm3=mol)  103 0.02355 0.02449 2975 0.028047 7.1501 0.07337 22.392 24.038
Compressibility factor, Z ¼ PV=RT — 0.0000082 0.9986 0.00379 0.9662 0.2879 0.9990 0.9992
Heats of fusion and vaporization (J=mol) 444.8 — 7761.4 6812.3 — 0 — —
Specific heat (J=(mol K))
Cg, at saturation 46.07 53.313 –108.7 54.14 –53.2 (very large) — —
CP, at constant pressure — 53.27 29.13 54.28 30.77 (very large) 29.33 29.40
CV at constant volume — 35.65 20.81 29.64 21.28 (38.7) 20.96 21.04
Specific heat ratio, g ¼ CP=CV — 1.494 1.400 1.832 1.446 (large) 1.40 1.40
Enthalpy (J=mol) 6634.46 6189.6 1571.8 4270.3 2542.0 1032.2 7937.8 8525.1
Internal energy (J=mol) 6634.4 6189.6 1120.0 4273.1 1817.5 662.3 5668.9 6089.5
Entropy (J=(mol K)) 58.92 67.11 209.54 94.17 169.68 134.42 202.4 204.5
Velocity of sound (m=s) — 1159 141 903 178 164 315 326
Viscosity
(N sec=m3)  103 — 0.6194 0.003914 0.1958 0.00685 (0.031) 0.01924 0.02036
Centipoises — 0.6194 0.003914 0.1958 0.00685 (0.031) 0.01924 0.02036
Thermal conductivity, k(mW=(cm K)) — 1.929 0.04826 1.515 0.08544 —b 0.2428 0.2575
Prandtl number, Npr ¼ uCp=k — 5.344 0.7392 2.193 0.7714 — 0.7259 0.7265
Dielectric constant, e pffiffiffiffi (1.614) 1.5687 1.000004 1.4870 1.00166 1.17082 1.00053 1.00049
Index of refraction, n ¼ ea (1.271) 1.2525 1.000002 1.219 1.00083 1.0820 1.00027 1.00025
Surface tension (N=m)  103 — 22.65 — 13.20 — 0 — —
Equiv. vol.=vol. liquid at NBP 0.8397 0.8732 106,068 1 254.9 2.616 798.4 857.1
a
Long wavelengths.
b
Anomalously large.
Gas constant: R ¼ 62,365.4 cm3 mmHg=(mol K).

97
Values in parentheses are estimates based on the NBS-1955 temperature scale using 90.180 K as a fixed point for the normal boiling temperature of oxygen. (The IPTS-1968
temperature scale uses 90.188 K as the normal boiling temperature of oxygen, but the reported data used in these tables have not yet been converted to the new temperature
scale.) 273.15 K ¼ 0 C ¼ 32 F ¼ 491.67 R base point (zero values) for enthalpy, internal energy, and entropy are 0 K for the ideal gas at 1 atm pressure.

© 2005 by Marcel Dekker


Table 3.10 Fixed Point Properties of Oxygen (English Units)

Triple point Normal boiling point Standard conditions

98
Property Solid Liquid Vapor Liquid Vapor Critical point STP(32 F) NTP(68 F)

Temperature ( F) 361.84 361.84 361.84 297.35 297.35 181.43 32.0 68.0


Pressure (psia) 0.0220 0.0220 0.0220 14.696 14.696 731.4 14.696 14.696
Density (lb=ft2) 84.82 81.57 0.0006715 71.23 0.2794 27.23 0.0892 0.0831
Specific volume (ft3=lb) 0.01179 0.01226 1489.2 0.01404 3.5793 0.03673 11.21 12.03
Compressibility factor, Z ¼ PV=RT — 0.0000082 0.9985 0.00379 0.9662 0.2879 0.9990 0.9992
Heats of fusion and vaporization (Btu=lb) 5.976 — 104.348 91.588 91.588 0 — —
Specific heat (Btu=lb  R)
Cs, at saturation 0.345 0.398 0.812 0.404 0.397 (very large) — —
CP, at constant pressure — 0.398 0.218 0.405 0.230 (very large) 0.219 0.220
CV, at constant volume — 0.266 0.155 0.221 0.159 (0.289) 0.157 0.157
Specific heat ratio, g ¼ CP=CV — 1.496 1.406 1.833 1.447 (large) 1.40 1.40
Enthalpy (Btu=lb) 89.192 83.216 21.132 57.412 34.176 13.88 106.72 114.62
Internal energy (Btu=lb) 89.192 83.216 15.057 57.450 24.435 8.90 76.22 81.875
Entropy (Btu=lb  R) 0.4401 0.50122 1.5651 0.70339 1.2674 1.004 1.5123 1.5278
Velocity of sound (ft=sec) — 3804 461 2963 583 537 1033 1070
Visosity, m
(lb=sec-ft)  105 — 41.62 0.263 13.16 0.460 (2.1) 1.293 1.368
(centipoise) — 0.6194 0.003914 0.1958 0.00685 (0.031) 0.01924 0.02036
Thermal conductivity (Btu=hr ft  R), k — 0.11156 0.00279 0.08758 0.00494 —b 0.01404 0.01489
Prandtl number, Npr ¼ mCP=k — 5.3437 0.7392 2.1929 0.7714 — 0.7259 0.7265
Dielectric constant, e pffiffiffiffi (1.614) 1.5687 1.000004 1.4870 1.00166 1.17082 1.00053 1.00049
Index of refraction, n ¼ n ¼ ea (1.1271) 1.2525 1.000002 1.219 1.00083 1.0820 1.00027 1.00025
3
Surface tension (lb=ft)  10 — 1.552 — 0.9046 — 0 — —
Equiv. vol.=vol. liquid at NBP 0.8397 0.8732 106,068 1 254.9 2.616 798.4 857.1
a
Long wavelengths.
b
Anomalously large.
Gas constant: R ¼ 0.335385 ft3 psi=(lb  R).
Values in parentheses are estimates based on the NBS-1955 temperature scale using 90.180 K as a fixed point for the normal boiling temperature of oxygen. (The IPTS-1968

Chapter 3
temperature scale used 90.188 K as the normal boiling temperature of oxygen, but the reported data used in these tables have not yet been converted to the new temperature
scale.) 273.15 K ¼ 0 C ¼ 32 F ¼ 491.67 R. Calculated from property values given in this table. Base point (zero values) for enthalpy, internal energy, and entropy are 0 K for the
ideal gas at 1 atm pressure.

© 2005 by Marcel Dekker


5367-4 Flynn Ch03 R2 090804

Cryogenic Fluids 99

Figure 3.14 Temperature–entropy chart for oxygen, English units.

Liquid oxygen equipment must also be designed of construction materials


incapable of initiating or sustaining a reaction. Only a few polymeric materials (plas-
tics), for example, can be used in the design of such equipment, as most will react
violently with oxygen under mechanical impact. Also, reactive metals such as
titanium or aluminum must be used cautiously, because they are potentially hazar-
dous. Once the reaction is started, for instance, an aluminum pipe containing oxygen
burns rapidly and intensely. With proper design and care, however, liquid oxygen
systems can be operated safely. For a complete discussion of this important topic,
please see the comprehensive work of Schmidt and Forney (1975).

© 2005 by Marcel Dekker


5367-4 Flynn Ch03 R2 090804

100 Chapter 3

Figure 3.15 Temperature–entropy chart for oxygen, metric units.

Oxygen is manufactured in large quantities by distillation of liquid air because


oxygen is the second most abundant substance in air (20.95% by volume or 23.2% by
weight). In comparison, the atmosphere of Mars contains about 0.15% oxygen.
Oxygen under excited conditions is responsible for the bright red and yellow-green
colors of the aurora borealis. The element and its compounds make up 49.2% by
weight of the earth’s crust.
Property data for oxygen are available from Roder and Weber (1972) unless
otherwise noted.

© 2005 by Marcel Dekker


Cryogenic Fluids

5367-4 Flynn Ch03 R2 090804


Figure 3.16 Mollier chart for oxygen.

101
© 2005 by Marcel Dekker
5367-4 Flynn Ch03 R2 090804

102 Chapter 3

Figure 3.17 Joule–Thomson inversion curve for oxygen.

c. Heat of Vaporization. The heat of vaporization (Fig. 3.18) is the


heat required to convert a unit mass of a substance from the liquid to the vapor
state at constant pressure. Its units are joules per mole or British thermal units
per pound:

Unit Triple point Boiling point

J=mol 7761 6812


Btu=lb 104.3 91.588

The uncertainty is estimated to be  10 J=mol (0.13 Btu=lb).


d. Heat of Sublimation. The heat of sublimation is the heat required to vapor-
ize a unit mass of solid.

© 2005 by Marcel Dekker


5367-4 Flynn Ch03 R2 090804

Cryogenic Fluids 103

Table 3.11 Joule–Thomson Inversion Curve for Oxygen

T (K) P (atm) Density (mol=cm3) DPa (atm)

125 15.60 0.02942 1.83


130 52.71 0.02907 1.98
135 86.45 0.02869 2.13
140 119.17 0.02835 2.32
145 150.87 0.02802 2.58
150 178.30 0.02765 2.61
155 204.56 0.02729 2.91
160 232.57 0.02700 2.92
165 256.29 0.02666 3.13
170 278.96 0.02633 3.23
175 300.60 0.02600 3.43
180 321.28 0.02569 3.49
185 340.06 0.02537 3.81
190 357.65 0.02504 3.80
a
Estimated uncertainty.

Range of values:

20.0–23.8 K 23.8–43.8 K 43.8–54.35 K


Units (gas–a-solid) (gas–b-solid) (gas–g-solid)

J=mol 9309–9265 9216–9131 8389–8207


Btu=lb 125.1–124.5 123.8–122.7 112.7–110.3

The uncertainty is estimated to be 10 J=mol (0.13 Btu=lb), with the uncertainty


in temperature as large as 0.1 K.
e. Vapor Pressure. The vapor pressure is the pressure, P(T), of a liquid and its
vapor in equilibrium. For P in atmospheres and T in kelvins, the equation is

ln P ¼ A1 þ A2 T þ A3 T 2 þ A4 T 3 þ A5 T 4 þ A6 T 5 þ A7 T 6 þ A8 T 7 ð3:1Þ

where

A1 ¼ 62:5967185 A5 ¼ 4:09349868  106


A2 ¼ 2:47450429 A6 ¼ 1:91471914  108
A3 ¼ 4:68973315  102 A7 ¼ 5:13113688  1011
A4 ¼ 5:48202337  104 A8 ¼ 6:0265693  1014

The range of values for the vapor pressure of oxygen is:

Unit Triple point Critical point

atm 0.0015 49.77


psia 0.022 731

© 2005 by Marcel Dekker


5367-4 Flynn Ch03 R2 090804

104 Chapter 3

Figure 3.18 Heat of vaporization of oxygen.

The estimated uncertainty is 0.02% from the normal boiling point to the critical
point. Below the normal boiling point the uncertainty increases, reaching about 3%
at the triple point. The equivalent uncertainty is 0.02 K in the range of 60–154.58 K.
Figures 3.19 and 3.20 are graphs of the vapor pressure of oxygen.
f. Virial Coefficients. The virial coefficients (Table 3.12) are usually defined
from the virial equation in density,

P ¼ RTr½1 þ BðTÞr þ CðTÞr2 þ    ð3:2Þ

The virial coefficients are functions of temperature only.


Two coefficients, B(T) and C(T), are adequate to describe the PVT surface
accurately up to a density of about one-half critical. The uncertainty for B varies
from  30 cm3=mol at the boiling point to  0.25 cm3=mol for temperatures greater
than 150 K; for C, from  10,000(cm3=mol)2 at the boiling point to  30(cm3=mol)2
above 150 K.
The Boyle point (B ¼ 0) for oxygen is at 405.88 K.
g. Dielectric Constant. The dielectric constant can be calculated from an
extension of the Clausius–Mossotti relationship

e1
¼ Ar þ Br2 þ Cr3 ð3:3Þ
eþ2

where A ¼ 0.12361, B ¼ 3.2  104, C ¼ 1.21  103, and r is in g=cm3. The equa-
tion is valid over the range 54.35–340 K, 0.2–340 atm, and will yield reasonable
values upon extrapolation.

© 2005 by Marcel Dekker


5367-4 Flynn Ch03 R2 090804

Cryogenic Fluids 105

Figure 3.19 Vapor pressure of oxygen in atm and mmHg vs. temperature in K.

The uncertainty in the term e  1 varies from 0.15% at low densities to less than
0.05% at high densities.
h. Other Properties. Tables 3.13–3.15 give values for the viscosity of gaseous
and liquid oxygen as well as the density of saturated liquid oxygen at various tem-
peratures. Figures 3.21–3.29 were compiled by the National Bureau of Standards,
Cryogenic Engineering Laboratory for the US Air Force, Wright Air Development
Division in WADD Technical Report 60–56, V. J. Johnson ed., October 1960. They
are commonly referred to as the WADD data. In Figs. 3.21–3.29 you will find best
values of the melting curve, dielectric constant, density, specific heats, and viscosity,
for oxygen.
3.1.2. Uses of Oxygen
Commercial oxygen demand in the United States is about 45  109 kg (50 million
tons) per year, at an average growth rate of 5.1% annually. For the most part,

© 2005 by Marcel Dekker


5367-4 Flynn Ch03 R2 090804

106 Chapter 3

Figure 3.20 Vapor pressure of oxygen in psia vs. temperature in  R.

tonnage oxygen (that delivered via pipeline) is consumed in steel plants. Present con-
sumption in other industries has seldom reached levels requiring individual oxygen
plants; those industries are usually supplied with liquid ‘‘merchant oxygen’’, which
is the name given to liquid oxygen delivered to the consumer by truck. The tonnage
oxygen plants currently fill about 80% of the oxygen demand.
The first continuous machine for liquefying air was built by Karl von Linde of
Munich, Germany, in 1895. The first liquid air installation of commercial size in the
United States was built for the American Linde Co., now Linde Division of Union
Carbide, in Niagara Falls, New York, in 1907. It had a capacity of 21,240 m3
(750,000 cu ft) per month of oxygen, equivalent to about 907 kg (1 ton) per day.
There is no substitute for oxygen in any of its uses, and it is not recycled or
reclaimed.
a. Steel. About 60–70% of the oxygen produced in the United States is
consumed in the manufacture of iron and steel. Concurrently with the development
of pelletizing and other improvements in blast furnace practice, low-cost tonnage

© 2005 by Marcel Dekker


5367-4 Flynn Ch03 R2 090804

Cryogenic Fluids 107

Table 3.12 Second and Third Virial Coefficients for Oxygen

T (K) B (cm3=mol) C(cm3=mol)2 T (K) B (cm3=mol) C (cm3=mol)2

85 267.78 21,462 210 44.66 1580


90 240.67 12,764 215 42.47 1537
95 217.51 7058 220 40.02 1498
100 197.54 3326 225 37.90 1461
105 180.20 904 230 35.89 1428
110 165.05 644 235 33.98 1397
115 151.71 1609 240 32.17 1368
120 139.91 2187 245 30.45 1342
125 129.41 2507 250 28.81 1317
130 120.02 2659 255 27.25 1294
135 111.59 2702 260 25.77 1273
140 103.98 2677 265 24.34 1253
145 97.08 2611 270 22.98 1234
150 90.81 2522 275 21.68 1217
155 85.09 2423 280 20.44 1201
160 79.84 2320 285 19.24 1186
165 75.02 2219 290 18.09 1172
170 70.58 2122 295 16.98 1160
175 66.48 2031 300 15.92 1149
180 62.67 1948
185 59.14 1871
190 55.85 1801
195 52.77 1738
200 49.89 1680
205 47.20 1628

oxygen became available in the 1950s and was instrumental in the development of
the basic oxygen furnace (BOF), which is a converter using oxygen instead of air.
b. Chemical. The chemical industry uses about 12% of the total oxygen
produced. Major uses are in partial oxidation of methane to produce acetylene; in
the oxidation of ethylene (the chief hydrocarbon produced in the United States) to

Table 3.13 Viscosity of Gaseous Oxygen at 1 atm

Temp. (K) Viscosity (cP)  103 Temp. (K) Viscosity (cP)  103

100 7.715 200 14.775


110 8.500 210 15.400
120 9.225 220 16.000
130 10.000 230 16.620
140 10.725 240 17.200
150 11.440 250 17.790
160 12.135 260 18.350
170 12.820 270 19.000
180 13.490 280 19.450
190 14.140 290 20.000
300 20.645

Source: Hilsenrath et al. (1955).

© 2005 by Marcel Dekker


5367-4 Flynn Ch03 R2 090804

108 Chapter 3

Table 3.14 Viscosity of Liquid Oxygen

Rudenko and Shubnikow (1934) Rudenko (1934)

Temp. (K) Viscosity (cP) Temp. (K) Viscosity (cP)

54.4 0.873 90 0.190


54.5 0.863 111 0.123
54.6 0.821 112 0.121
54.9 0.772
56.4 0.717 125.8 0.110
57.1 0.638 138.2 0.100
57.4 0.648
59.7 0.631 145.5 0.098
61.7 0.521 154.1 0.090
63.5 0.476
65.4 0.435
68.9 0.377
72.3 0.323
77.4 0.273
80.0 0.250
90.1 0.190

Source: Rudenko (1939), Rudenko and Shubnikow (1934).

produce ethylene oxide, an important chemical intermediate; and in various pro-


cesses to produce hydrogen from hydrocarbons for the manufacture of chemicals
such as ammonia.
c. Nonferrous Metals. About 6% of the oxygen produced is used by the
nonferrous metals industry. It is used to enrich air in primary and secondary lead
blast furnaces and in copper reverberatory furnaces, converters, flash smelters,
and autogenous and continuous copper smelting processes.
Example 3-1 (Oxygen Sample Problem). What is the boiling point of oxygen at
150 psia? What is the heat of vaporization at this pressure?

Table 3.15 Density of Saturated Liquid Oxygen

Van Itterbeek (1955) Mathias and Onnes (1911)

Temp. (K) Density (g=cm3) Temp. (K) Density (g=cm3)

61 1.282 62.7 1.2746


65 1.263 91.1 1.1415
70 1.239 118.6 0.9758
75 1.215 132.9 0.8742
80 1.191 143.2 0.7781
85 1.167 149.8 0.6779
90 1.142 152.7 0.6032
154.3a 0.4299
a
Critical point (current accepted value is 154.77 K based on present international scale of 0 C ¼ 273.15 K
instead of the Leiden temperature scale of 0 C ¼ 273.09 K).
Source: Van Itterbeek (1955), Mathias and Onnes (1911).

© 2005 by Marcel Dekker


5367-4 Flynn Ch03 R2 090804

Cryogenic Fluids 109

Figure 3.21 Melting curve for oxygen (NASA SP3071).

The T–S diagram for oxygen (Fig. 3.30) contains much more information than
first meets the eye. The answers to the above questions are readily found on the
oxygen T–S chart.
First, find the 150 psia isobar for oxygen and follow it to the two–phase region.
The isobars and isotherms are coincident horizontal lines under the two–phase
dome. Therefore, the boiling point of oxygen is read directly from the temperature
scale to be about 216 R. The published value is 216.38 R at 150 psia.
Likewise, the heat of vaporization can be found in two independent ways.
First, it is only necessary to read the enthalpy values for saturated vapor (Hg) and
saturated liquid (Hf). Along the 150 psia isobar we find Hg ¼ 40 Btu=lbm and
Hf ¼ –34 Btu=lbm. The difference is Hfg ¼ Hg  Hf ¼ 40  (34) ¼ 74 Btu=lbm. The
published value is 74.39 Btu=lbm.
Heat of vaporization is also TDS. From the TS chart we find Sg ¼
1.167 Btu=(lbm  R) and Sf ¼ 0.824 Btu=(lbm  R). Accordingly, DS ¼ 1.167  0.824 ¼
0.343 Btu=(lbm  R). TDS in appropriate units is (216.4 R)  [0.343 Btu=
(lbm  R)] ¼ 74.2 Btu=lbm, which compares well with the published value of
74.4 Btu=lbm.

3.2. Nitrogen
Nitrogen (mol wt. 28.0134) has two stable isotopes of mass numbers 14 and 15 with a
relative abundance of 10,000:38. Liquid nitrogen is of considerable importance to the
cryogenic engineer because it is a safe refrigerant. Because it is rather inactive

© 2005 by Marcel Dekker


5367-4 Flynn Ch03 R2 090804

110 Chapter 3

Figure 3.22 Dielectric constant of oxygen (NASA SP3071).

chemically and is neither explosive nor toxic, liquid nitrogen is commonly used in
hydrogen and helium liquefaction cycles as a precoolant.
Nitrogen is the major constituent of air (78.09% by volume or 75.45% by
weight). The atmosphere of Mars, by comparison, is 2.6% nitrogen. The estimated
amount of this element in out atmosphere is more than 4000 billion tons. From this
inexhaustible source, it can be obtained by liquefaction and fractional distillation.
Liquid nitrogen is a clear, colorless fluid that resembles water in appearance.
At 1 bar pressure, liquid nitrogen boils at 77.3 K (139.2 R) and freezes at 63.2 K
(113.8 R). Saturated liquid nitrogen at the normal boiling point has a density of
808.9 kg=m3 (50.5 lbm=ft3) in comparison with water at 520 R (60 F), which has a
density of 998.0 kg=m3 (62.3 lbm=ft3). One of the significant differences between
the properties of liquid nitrogen and those of water (apart from the differences in
normal boiling points) is that the heat of vaporization of nitrogen (Table 3.16) is
more than an order of magnitude less than that of water. At its normal boiling point,
liquid nitrogen has a heat of vaporization of 198.3 kJ=kg (85.32 Btu=lbm), while
water has a heat of vaporization of 2255 kJ=kg (970.3 Btu=lbm).
Table 3.17 gives the physical constants of nitrogen.

© 2005 by Marcel Dekker


5367-4 Flynn Ch03 R2 090804

Cryogenic Fluids 111

Figure 3.23 Dielectric constant of oxygen, extended range (NASA SP3071).

Figure 3.24 Density of oxygen (NASA SP3071).

© 2005 by Marcel Dekker


5367-4 Flynn Ch03 R2 090804

112 Chapter 3

Figure 3.25 Specific heat of oxygen (NASA SP3071).

Table 3.18 gives some of the important fixed points for nitrogen.
Figures 3.31–3.33 are TS charts for nitrogen. Figure 3.34 is the Joule–
Thomson inversion curve.
Nitrogen has a potential safety hazard in that a bare pipe of liquid nitrogen at
77 K will condense an air mixture containing approximately 50% liquid oxygen.
Thus, one must exercise extreme caution that so-called inert liquid nitrogen does not
in fact become an unsuspected source of liquid oxygen, with the attendant risks cited
earlier. Several explosions and deaths have been attributed to the phenomenon of
oxygen enrichment of the atmosphere in the presence of liquid nitrogen cooled
surfaces. See Chapter 10.

3.2.1. Properties of Nitrogen


Property data are from Jacobsen et al. (1973) unless otherwise noted.
a. Heat of Vaporization.

Units Triple point (63.148 K) Boiling point (77.313 K)

J=g 215.1 198.8


J=mol 6025.7 5569.1
Btu=lb 92.57 85.55
Btu=(lb mol) 2593.2 2396.5

© 2005 by Marcel Dekker


5367-4 Flynn Ch03 R2 090804

Cryogenic Fluids 113

Figure 3.26 Specific heat (CP) of solid oxygen (NASA SP3071).

b. Vapor Pressure. For solid nitrogen.

359:093
log p ¼ a  ð3:4Þ
T

where T is in K, a ¼ 7.65894 for p in mmHg, a ¼ 4.77813 for p in atm and a ¼ 5.94532


for p in psia.
or, for p below 760 mmHg,

334:64
log p ¼  þ 7:577  0:00476 T ð3:5Þ
T

© 2005 by Marcel Dekker


5367-4 Flynn Ch03 R2 090804

114 Chapter 3

Figure 3.27 Specific heat (CP) of solid oxygen below 4 K (NASA SP3071).

where p is in mmHg and T is in K. (This equation is also valid for liquid nitrogen
from 63.14 to 77.35 K; 96.4–760 mmHg.)
For liquid nitrogen,

255:821
log p ¼ 6:49594  ð3:6Þ
T  6:600
where p is in mm Hg and T is in K for the range 64–78 K
or

316:824
log p ¼  þ 4:47582  0:0071701T þ 2:940  105 T 2 ð3:7Þ
T
where, p is in atm, T is in K and 1–32 atm.

© 2005 by Marcel Dekker


5367-4 Flynn Ch03 R2 090804

Cryogenic Fluids 115

Figure 3.28 Specific heat of oxygen at constant pressure (NASA SP3071).

Figure 3.29 Viscosity of oxygen.

© 2005 by Marcel Dekker


5367-4 Flynn Ch03 R2 090804

116 Chapter 3

Figure 3.30 Temperature–entropy chart for oxygen, sample problem.

c. Dielectric Constant. The dielectric constant of nitrogen may be calculated


from the Clausius–Mossotti equation,
 
e1 1
¼p ð3:8Þ
eþ2 r

© 2005 by Marcel Dekker


5367-4 Flynn Ch03 R2 090804

Cryogenic Fluids 117

Table 3.16 Heat of Vaporization of Nitrogen

Heat of vaporization, hv

Temp. (K) J=mol cal=mol

Solid nitrogen
62.00 6775.0
62.0018 6787.4
62.0172 6762.4
Liquid nitrogen
67.9588 5901.6
67.9620 5899.0
68.00 5899.0
73.0913 5739.1
73.0887 5732.1
73.10 5735.2
77.395 5735.2
78.00 5579.4
78.0147 5563.1
78.0153 5571.8
80 1313
85 1266
90 1213
95 1155
100 1086
105 1010
110 918
115 803
120 643
125 328
126.1a 0
a
Critical point.
Source: Furukawa and McCoskey (1953), Millar and Sullivan
(1928).

where e is the dielectric constant, r is the density, and p is the specific polarization, a
property of the substance having dimensions of specific volume:

p ¼ An þ Bn r þ Cn r2 ð3:9Þ

where p is the specific polarization in cm3=mol and r is in units of mol=cm3.


The parameters An, Bn, and Cn are

An ¼ 4:389 cm3 =mol; Bn ¼ 2:2ðcm3 =molÞ2 ; Cn ¼ 114:0ðcm3 =molÞ3

d. Other Properties. Tables 3.19–3.23 present data on the solid and liquid
vapor pressure, density, viscosity of nitrogen. Figures 3.35–3.53, taken from the
WADD report mentioned before, show the heat of vaporization, vapor pressure,
melting curve, density, thermal expansivity, thermal conductivity, specific heats,
dielectric constant, surface tension, and viscosity of nitrogen.

© 2005 by Marcel Dekker


5367-4 Flynn Ch03 R2 090804

118 Chapter 3

Table 3.17 Nitrogen Physical Constants

Nitrogen
Chemical Symbol: N2
Synonyms: LIN (liquid only)
CAS Registry Number: 7727-37-9
DOT Classification: Nonflammable gas
DOT Label: Nonflammable gas
Transport Canada Classification: 2.2
UN Number: UN 1066 (compressed gas); UN 1977 (refrigerated liquid)

US Units SI Units

Physical Constants
Chemical formula N2 N2
Molecular weight 28.01 28.01
Density of the gas at 70 F (21.1 C) and 1 atm 0.072 lb=ft3 1.153 kg=m3
Specific gravity of the gas at 70 F (21.1 C) 0.967 0.967
and 1 atm (air ¼ 1)
Specific volume of the gas at 70 F (21.1 C) 13.89 ft3=lb 0.867 m3=kg
and 1 atm
Density of the liquid at boiling point and 1 atm 50.47 lb=ft3 808.5 kg=m3
Boiling point at 1 atm 320.4 F 195.8 C
Melting point at 1 atm 345.8 F 209.9 C
Critical temperature 232.4 F 146.9 C
Critical pressure 493 psia 3399 kPa abs
Critical density 19.60 lb=ft3 314.9 kg=m3
Triple point at 1.81 psia (12.5 kPa abs) 346.0 F 210.0 C
Latent heat of vaporization at boiling point 85.6 Btu=lb 199.1 kJ=kg
Latent heat of fusion at melting point 11.1 Btu=lb 25.1 kJ=kg
Specific heat of the gas at 70 F (21.1 C)
and 1 atm
CP 0.249 Btu=(lb)( F) 1.04 kJ=(kg)( C)
CV 0.177 Btu=(lb)( F) 0.741 kJ=(kg)( C)
Ratio of specific heats, CP=CV 1.41 1.41
Solubility in water vol=vol at 32 F (0 C) 0.023 0.023
Weight of liquid at boiling point 6.747 lb=gal 808.5 kg=m3
Gas=liquid ratio (gas at 70 F (21.1 C) and 696.5 696.5
1 atm, liquid at boiling point, vol=vol)

3.2.2. Uses of Nitrogen


Nitrogen is used in two distinct ways: (1) as an element, which may be used either as
a gas to exclude air from industrial processes or as a liquid to provide refrigeration,
and (2) as fixed nitrogen, in compounds with other elements, in which form it is an
essential plant nutrient and a constituent of many important chemical products.
Elemental nitrogen is produced by cryogenic means and is used primarily for
two purposes. Most is used as a gas to provide an inert atmosphere in a variety of
chemical and metallurgical processes, electronic devices, and packages; about 10%
is used as a liquid to provide refrigeration. The major consuming areas are discussed
below.
a. Chemicals. Numerous chemical reactions require an inert atmosphere
to prevent oxidation, fire, or explosion, for drying or purging equipment, for

© 2005 by Marcel Dekker


5367-4 Flynn Ch03 R2 090804

Cryogenic Fluids 119

Table 3.18 Fixed Points for Nitrogen

Critical point
T ¼ 126.20 K (227.16 R)
P ¼ 33.55 atm (493.0 psia;3.399 MPa)
r ¼ 0.01121 mol=cm3 (19.60 lbm=ft3; 314 kg=m3)
Normal boiling point
T ¼ 77.347 K (139.225 R)
P ¼ 1 atm (14.696 psia; 101.3 kPa)
rgas ¼ 0.0001647 g mol=cm3 (0.2880 lbm=ft3; 4.614 kg=m3)
rliquid ¼ 0.02887 g mol=cm3 (50.49 lbm=ft3; 808.9 kg=m3)
Triple point
T ¼ 63.148 K (113.666 R)
P ¼ 0.1237 atm (1.818 psia; 12.53 kPa)
rgas ¼ 2.41  105 g mol=cm3 (4.21  102 lbm=ft3; 674 g=m3)
rliquid ¼ 0.03098 gmol=cm3 (54.18 lbm=ft3; 867.96 kg=m3)

Source: Jacobsen et al. (1973).

pressurizing, or as a dilutent to control gas reaction rates. About 25% of elemental


nitrogen is used to provide blanketing atmospheres for chemical processes.
b. Electronic Components. Gaseous nitrogen is used to pressurize telephone
and electric cables and to flush air out of and dry electronic tubes. Liquid nitrogen
is used as a coolant for lasers, masers, and infrared detectors. Use in electronic com-
ponents constitutes about 15% of the demand for elemental nitrogen.
c. Food Products. The food industry is the largest consumer of liquid nitro-
gen. Evaporation of liquid nitrogen is a fast, uncomplicated, reliable, and inexpen-
sive method of providing refrigeration. Although it is not competitive for small
installations requiring moderately low temperatures, such as household refrigerators,
it provides a much lower temperature than competitive large refrigeration units and
is particularly useful for fast freezing, as in the meat backing industry. Liquid nitro-
gen is especially convenient for refrigeration during transportation of foodstuffs
because it is relatively simple to install compared to mechanical refrigerators and
has the additional advantage of naturally providing an inert atmosphere. This use
is slowing down in its rate of growth and now accounts for about 5% of the nitrogen
market.
d. Iron and Steel. A large use of elemental nitrogen, about 15% of the total, is
in annealing stainless steel; the nitrogen may be mixed with hydrogen, which acts as a
deoxidizing agent. Annealing and heat treating of other steel mill products such as
tinplate also are conducted in a nitrogen atmosphere.
e. Nonferrous Metal Production. Nitrogen gas is used for preparing reactive
metals, such as the rare earths and sodium and potassium, and to degas molten
aluminum. Liquid nitrogen is used in shrink-fitting metals and in the grinding of
heat-sensitive materials. About 2% of elemental nitrogen production goes into all
these uses.
f. Other Applications. Nitrogen is used in various aerospace applications
(about 5% of the market) such as for transferring propellants and in wind tunnels.
Nitrogen appears to have developed its own big new use over the past few years—
enhanced oil recovery (EOR), used to boost the output from old oilfields. This use
now accounts for 15% of all nitrogen consumed in the United States.

© 2005 by Marcel Dekker


5367-4 Flynn Ch03 R2 090804

120 Chapter 3

Figure 3.31 Temperature–entropy chart for nitrogen.

The diversity of nitrogen markets resulted from producers’ efforts over


many years to find some way to sell this formerly discarded coproduct of many
air separation plants. Oddly, producers succeeded so well that they now have the

© 2005 by Marcel Dekker


5367-4 Flynn Ch03 R2 090804

Cryogenic Fluids 121

Figure 3.32 Temperature–entropy chart for nitrogen.

reverse problem at many plants; they are looking for new oxygen uses to balance the
demand for nitrogen.
Example 3–2 Nitrogen Sample Problem. Nitrogen gas at 138 K and 220 atm
pressure is expanded through a throttling valve. What is the final temperature?
What is the fraction liquefied?
This problem statement is typical of the conditions found in a Joule–Thomson
nitrogen liquefier. The nitrogen gas has been compressed to 200 atm and precooled
to 138 K in a countercurrent heat exchanger just prior to expansion through the J–T

© 2005 by Marcel Dekker


5367-4 Flynn Ch03 R2 090804

122 Chapter 3

Figure 3.33 Temperature–entropy chart for nitrogen.

valve. We know the starting conditions, 138 K, 200 atm. The thermodynamic process
is expansion through a J–T valve, which is isenthalpic.
The nitrogen T–S diagram of Fig. 3.54 contains all the information we need to
solve this problem.
The starting point is 138 K, 200 atm, which also fixes the enthalpy at 160 J=gK.
Since the process is isenthalpic, we follow the 160 J=gK isenthalp to the final pressure
of 1 atm. Note that temperature and pressure lines are coincident and horizontal in
the two-phase liquid–vapor region. The final temperature is easily found, namely the
NBP of nitrogen, 77 K. The fraction liquefied is given by the lines of constant quality
(X) in the two-phase region. Our final quality falls between X ¼ 60% vapor and
X ¼ 70% vapor. We estimate the final quality to be about 63% vapor.

© 2005 by Marcel Dekker


5367-4 Flynn Ch03 R2 090804

Cryogenic Fluids 123

Figure 3.34 Joule–Thomson inversion curve for nitrogen.

Table 3.19 Vapor Pressure of Solid and Liquid Nitrogen

Vapor Vapor Vapor


Temp. (K) pressure (atm) Temp. (K) pressure (atm) Temp.(K) pressure (atm)

52 0.0075 76 0.8461 104 10.07


54 0.0134 77.35  .02b 1.000 105.0 10.71
56 0.0232 78 1.073 106 11.42
57.9 0.0379 80.0 1.36 108 12.91
58 0.0387 80 1.341 110.0 14.54
59.0 0.0496 82 1.657 110 14.52
60 0.0621 84 2.026 112 16.26
61.0 0.0787 85.0 2.25 114 18.15
62 0.0968 86 2.460 115.0 19.28
63.14a 0.1268 88 2.967 116 20.20
63.156 0.1237 90.0 3.54 118 22.41
64 0.1439 90 3.548 120.0 25.04
64.55 0.1591 92 4.203 120 24.81
66 0.2028 94 4.937 122 27.40
68 0.2797 95.0 5.31 124 30.21
68.4 0.3005 96 5.76 125.0 31.94
70 0.3784 98 6.68 126 33.27
72 0.5033 100.0 7.67 126.1c 33.49
74 0.6579 100 7.70 126.1  0.1c 33.5
74.9 0.7386 102 8.83
a
Melting point.
b
Normal boiling point.
c
Critical point.
Source: Hoge (1950); Hoge and King (1950); Keesom (1922, 1923); Mathias and Crommelin (1924).

© 2005 by Marcel Dekker


5367-4 Flynn Ch03 R2 090804

124 Chapter 3

Table 3.20 Density of Saturated Liquid Nitrogen

Temperature (K) Density (g=cm3)

63.14a
64.73 0.8622
77.31 0.8084
77.32b
77.5
78.00 0.8043
90.58 0.7433
99.36 0.6922
111.89 0.6071
119.44 0.5332
125.01 0.4314
125.96c 0.31096

Note: The Leiden temperature scale (0 C ¼ 273.09 K) was used.


a
Triple point temperature.
b
Normal boiling temperature.
c
Critical temperature.
Source: Gerold (1921); Mathias and Crommelin (1924).

3.3. Air
Dry air (avg mol wt. approximately 29.00) is a mixture consisting principally of
nitrogen, oxygen, and argon with traces of other gases as shown in Table 3.25.
When air is liquefied, the carbon dioxide is usually removed; thus, for practical
purposes, liquid air can be considered to consist of 78% nitrogen, 21% oxygen, and
1% argon, the other constituents being present in negligible amounts. Sometimes the
presence of argon is ignored and liquid air is considered to be a binary mixture of
21% oxygen and 79% nitrogen. Since argon has a vapor pressure between those
of oxygen and nitrogen, this assumption is a rather good approximation for some
purposes.
Liquid air has a density of 874 g=L at its normal boiling point of 78.9 K.
Figure 3.55 is a T–S diagram for air.

Table 3.21 Viscosity of Liquid Nitrogen

Temp. (K) Viscosity (cP)

63.9 0.292
64.3 0.290
64.8 0.284
69.1 0.231
69.25 0.228
71.4 0.209
76.1 0.165
77.33 0.158
111.7 0.074

Source: Rudenko (1939); Rudenko and Shubnikow (1934).

© 2005 by Marcel Dekker


5367-4 Flynn Ch03 R2 090804

Cryogenic Fluids 125

Table 3.22 Viscosity of Gaseous Nitrogen


above 1 atm at T ¼ 298 K

Pressure (atm) Viscosity (cP)  102

1 1.778
2 1.780
5 1.784
10 1.793
20 1.809
30 1.827
40 1.846
50 1.867
60 1.890
70 1.916
80 1.946a
90 1.978a
100 2.015a
a
Extrapolated.
Source: Kestin and Wang (1958).

3.3.1. Properties of Air


All data are from Johnson (1960) and Din (1956) unless otherwise noted.
Table 3.24 gives the physical constants for air.
a. Heat of Vaporization. The heat of vaporization is usually defined for a sin-
gle component and is the heat required to completely change a given quantity of
liquid at its bubble point (boiling point) to vapor at its dew point at constant tem-
perature and pressure. However, for air, which is a multicomponent system, only
the pressure remains constant while the temperature rises throughout the process.
The data given in Table 3.26 are for air at the temperature and pressure of
the saturated liquid (bubble point). Table 3.27 presents data on the latent heat of
vaporization of an oxygen–nitrogen mixture at 1 atm.

Table 3.23 Viscosity of Gaseous Nitrogen at 1 atm

Temp. (K) Viscosity (cP) Temp. (K) Viscosity (cP)

90 0.006 298 210 0.013 499


100 0.006 975 220 0.014 029
110 0.007 631 230 0.014 547
120 0.008 264 240 0.015 052
130 0.008 876 250 0.015 547
140 0.009 484 260 0.016 031
150 0.010 083 270 0.016 502
160 0.010 676 280 0.016 960
170 0.011 253 290 0.017 410
180 0.011 829 298.1 0.017 777
190 0.012 394 300 0.017 857
200 0.012 954

Source: Johnston and McCloskey (1940).

© 2005 by Marcel Dekker


5367-4 Flynn Ch03 R2 090804

126 Chapter 3

Figure 3.35 Heat of vaporization of nitrogen.

b. Vapor Pressure. It should be noted that there are separate curves for the
vapor pressure of saturated liquid (bubble) and saturated vapor (dew). Air is a multi-
component mixture, and at a given pressure the saturated liquid is at a different tem-
perature than the saturated vapor (see Table 3.28 and Figure 3.56).

© 2005 by Marcel Dekker


5367-4 Flynn Ch03 R2 090804

Cryogenic Fluids 127

Figure 3.36 Vapor pressure of solid and liquid nitrogen (entire pressure range).

Bubble curve:
290:70
log P ¼ 3:5713  þ 0:001494T ð3:10Þ
T

© 2005 by Marcel Dekker


5367-4 Flynn Ch03 R2 090804

128 Chapter 3

Figure 3.37 Vapor pressure of solid and liquid nitrogen (low pressure range).

Dew curve:
333:88
log P ¼ 4:0816  ð3:11Þ
T
where P is in atm and T is kelvin.

© 2005 by Marcel Dekker


5367-4 Flynn Ch03 R2 090804

Cryogenic Fluids 129

Figure 3.38 Melting curve for nitrogen.

© 2005 by Marcel Dekker


5367-4 Flynn Ch03 R2 090804

130 Chapter 3

Figure 3.39 Density of saturated liquid nitrogen.

© 2005 by Marcel Dekker


5367-4 Flynn Ch03 R2 090804

Cryogenic Fluids 131

Figure 3.40 Density of gaseous nitrogen (saturated vapor) to the critical point.

c. Thermal Conductivity. See Table 3.29


d. Density. See Table 3.30
e. Vapor–Liquid Equilibrium Constants for the Major Components of Air. The
relationship between the mole fraction composition of a component (i) in the liquid

© 2005 by Marcel Dekker


5367-4 Flynn Ch03 R2 090804

132 Chapter 3

Figure 3.41 Density of gaseous nitrogen (saturated vapor) triple point to the NBP.

phase (xi) and the vapor phase (yi) at equilibrium is given by the equilibrium
constant (Ki) :
Ki ¼ yi =xi ð3:12Þ

© 2005 by Marcel Dekker


5367-4 Flynn Ch03 R2 090804

Cryogenic Fluids 133

Figure 3.42 Expansivity of liquid nitrogen along the melting curve.

If both the liquid and vapor phases are ideal mixtures, it can be shown from
Raoult’s and Dalton’s laws that

Ki ¼ Pi =P ð3:13Þ

where Pi is the vapor pressure of component i and P are total system pressure.

© 2005 by Marcel Dekker


5367-4 Flynn Ch03 R2 090804

134 Chapter 3

Figure 3.43 Thermal conductivity of saturated liquid nitrogen.

For real systems, values of K are determined experimentally and are often
tabulated as the distribution coefficient function, ln (Kp=po), where po is some refer-
ence pressure, such as 101.3 kPa (1 atm). Values for the distribution coefficient func-
tion for nitrogen, oxygen, and argon are given in Table 3.31.

© 2005 by Marcel Dekker


Cryogenic Fluids

5367-4 Flynn Ch03 R2 090804


135
Figure 3.44 Thermal conductivity of nitrogen.

© 2005 by Marcel Dekker


5367-4 Flynn Ch03 R2 090804

136 Chapter 3

Figure 3.45 Specific heat (CP) of solid nitrogen.

© 2005 by Marcel Dekker


5367-4 Flynn Ch03 R2 090804

Cryogenic Fluids 137

Figure 3.46 Specific heat (CV) of gaseous nitrogen.

© 2005 by Marcel Dekker


5367-4 Flynn Ch03 R2 090804

138 Chapter 3

Figure 3.47 Specific heat (CP) of saturated liquid nitrogen.

© 2005 by Marcel Dekker


5367-4 Flynn Ch03 R2 090804

Cryogenic Fluids 139

Figure 3.48 Specific heat (CP) of gaseous nitrogen.

© 2005 by Marcel Dekker


5367-4 Flynn Ch03 R2 090804

140 Chapter 3

Figure 3.49 Dielectric constant for solid and liquid nitrogen at 1 atm.

© 2005 by Marcel Dekker


5367-4 Flynn Ch03 R2 090804

Cryogenic Fluids 141

Figure 3.50 Surface tension of saturated liquid nitrogen.

© 2005 by Marcel Dekker


5367-4 Flynn Ch03 R2 090804

142 Chapter 3

Figure 3.51 Viscosity of liquid nitrogen.

The composition of the liquid and vapor phases of a two-component mixture


may be determined from values of the distribution coefficients for each component.
From Eq. (3.12),
y1 ¼ K1 x1
ð3:14Þ
y2 ¼ K2 x2 ¼ K2 ð1  x1 Þ
For a two-component mixture.
y1 þ y2 ¼ 1 ¼ K1 x1 þ K2 ð1  x1 Þ ð3:15Þ

Solving for the mole fraction of component 1 in the liquid phase,


1  K2
x1 ¼ ð3:16Þ
K1  K2
The mole fraction of component 1 in the vapor phase is found by combining Eq.
(3.13 ) and (3.16 ) :
K1 ð1  K2 Þ
y1 ¼ k1 x1 ¼ ð3:17Þ
K1  K2

3.3.2. Uses of Liquid Air


Liquid air was once widely used as a refrigerant for low-temperature investigations,
but the relative simplicity of producing liquid nitrogen from liquid air by distillation

© 2005 by Marcel Dekker


5367-4 Flynn Ch03 R2 090804

Cryogenic Fluids 143

Figure 3.52 Viscosity of gaseous nitrogen at 1 atm.

© 2005 by Marcel Dekker


5367-4 Flynn Ch03 R2 090804

144 Chapter 3

Figure 3.53 Viscosity of gaseous nitrogen above 1 atm.

has led to the gradual disuse of air. As it is essentially a mixture of nitrogen and oxy-
gen, air does not have the advantage of an invariant boiling temperature. A mixture
containing 80% nitrogen begins boiling at about 79 K, but since the initial vapor has
a composition richer than 80% nitrogen, the remaining liquid gradually becomes
richer in oxygen as boiling proceeds. This increases both the boiling temperature
of the remaining liquid and the hazards associated with oxygen-rich liquid.
This has caused explosions in the vacuum pumps used to reduce the pressure above
liquid air.
Most liquid air is now produced as an intermediate step in the production of
oxygen and nitrogen by distillation. The principal interest in liquid air is in the
preparation of pure nitrogen, oxygen, and rare gases.

3.4. Argon
Argon (at wt. 39.948) has three stable isotopes of mass numbers 36, 38, and 40,
which occur in a relative abundance in the atmosphere of 338:63:100,000.
Liquid argon is a clear, colorless fluid with properties similar to those of liquid
nitrogen. At 1 bar pressure, liquid argon boils at 87.3 K (157.1 R) and freezes
at 83.8 K (151.0 R). Saturated liquid argon at 0.987 bar (1 atm) is more dense than
oxygen, as one would expect, since argon has a larger molecular weight than oxygen
[argon density ¼ 1403 kg=m3 (87.5 lbm=ft3) for saturated liquid at 0.987 bar (1 atm) ].
See Fig. 3.57 for a T–S diagram of argon.
Argon is present in atmospheric air in a concentration of 0.934% by volume
or 1.25% by weight. The atmosphere of Mars contains 1.6% 40Ar and 5 ppm 36Ar,
for comparison. Since the boiling point of argon lies between that of liquid oxygen
and that of liquid nitrogen (slightly closer to that of liquid oxygen), a crude grade

© 2005 by Marcel Dekker


5367-4 Flynn Ch03 R2 090804

Cryogenic Fluids 145

Figure 3.54 Temperature–entropy chart for nitrogen, sample problem.

of argon (90–95% pure) can be obtained by adding a small auxiliary argon recovery
column in an air separation plant. About 900 kg (1 ton) of argon is recovered for
every 36  103 kg (40 tons) of oxygen. This is a little less than 50% of the argon
passing through the oxygen column and represents the point of present maximum

© 2005 by Marcel Dekker


5367-4 Flynn Ch03 R2 090804

146 Chapter 3

Table 3.24 Physical Constants for Air

Air
Synonyms: Compressed air, atmospheric air, the atmosphere (of the earth)
CAS Registry Number: None (for nitrogen, 7727–37-9; for oxygen, 7782–44-7)
DOT Classification: Nonflammable gas
DOT Label: Nonflammable gas
Transport Canada Classification: 2.2
UN Number: UN 1002 (compressed gas); UN 1003 (refrigerated liquid)

U.S. Units SI Units

Physical constants
Chemical name Air Air
Molecular weight 28.975 28.975
Density of the gas at 70 F (21.1 C) 0.07493 lb=ft3 1.2000 kg=m3
and 1 atm
Specific gravity of the gas at 70 F (21.1 C) 1.00 1.00
and 1 atm (air ¼ 1)
Specific volume of the gas at 70 F (21.1 C) 13.346 ft3=lb 0.8333 m3=kg
and 1 atm
Boiling point at 1 atm 317.8 F 194.3 C
Freezing point at 1 atm 357.2 F 216.2 C
Critical temperature 221.1 F 140.6 C
Critical pressure 547 psia 3771 kPa abs
Critical density 21.9 lb=ft3 351 kg=m3
Latent heat of vaporization at 88.2 Btu=lb 205 kJ=kg
normal boiling
point
Specific heat of gas at 70 F (21.1 C)
and 1 atm
CP 0.241 Btu=(lb)( F) 1.01 kJ=(kg)( C)
CV 0.172 Btu=(lb)( F) 0.720 kJ=(kg)( C)
Ratio of specific heats (CP=CV) 1.40 1.40
Solubility in water, vol=vol at 32 F (0 C) 0.0292 0.0292
Weight of liquid at normal boiling point 7.29 lb=gal 874 kg=m3
Density of liquid at boiling point 54.56 lb=ft3 874.0 kg=m3
and 1 atm
Gas=liquid ratio (liquid at boiling point, 728.1 728.1
gas at 70 F and 1 atm), vol=vol
Thermal conductivity
at 148 F (100 C) 0.0095 Btu=(hr) 0.0164 W=(m)( C)
(ft)( F=ft)
at 32 F (0 C) 0.0140 Btu=(hr) 0.0242 W=(m)( C)
(ft) ( F=ft)
at 212 F (100 C) 0.0183 Btu=(hr) 0.0317 W=(m)( C)
(ft) ( F=ft)

economic recovery. Some oxygen plants do not recover argon, and US argon
production is thus about a quarter of the theoretical maximum.
The first commercial production, amounting to about 227 m3 (8000 ft3), or
363 kg (800 lb), was carried out by the American Cyanamid Co., Niagara Falls,
Ontario, Canada, in 1914–1915. The argon was produced, not by the air liquefaction

© 2005 by Marcel Dekker


5367-4 Flynn Ch03 R2 090804

Cryogenic Fluids 147

Table 3.25 Composition of the Air Near Ground Level

Gas Mol wt. Percent by volume

Nitrogen 28 78.09
Oxygen 32 20.95
Argon 40 0.93
Carbon dioxide 44 0.02–0.04
Neon 20.2 18  10-4
Helium 4 5.3  104
Krypton 83.7 1.1  104
Hydrogen 2 0.5  104
Xenon 131.3 0.08  104
Ozone 48 0.02  104
Radon 222 7  1018

Figure 3.55 Temperature–entropy diagram for air.

© 2005 by Marcel Dekker


5367-4 Flynn Ch03 R2 090804

148 Chapter 3

Table 3.26 Heat of Vaporization of Air at the Bubble Point

Temp. (sat. liq.) Heat of vaporization



Pressure (atm) K R J=g Btu=lb

1 78.8 142.0 205.2 88.22


2 85.55 154.0 197.4 84.87
3 90.94 163.7 191.4 80.86
5 96.38 173.5 183.9 79.06
7 101.04 181.87 175.0 75.24
10 106.47 191.65 163.5 70.29
15 113.35 204.03 143.5 61.69
20 118.77 213.79 124.1 53.35
25 123.30 221.94 103.2 44.37
30 127.26 229.07 80.39 34.56
35 130.91 235.64 48.45 20.83
37.17 132.52 238.50 0 0

route that is now universal, but by chemical reaction of the other constituents of the
air with solids. Production of argon by liquefaction of air was begun by the Amer-
ican Linde Co. (now Linde Division of Union Carbide Corp.) in 1915. Production
was small until about 1943. Before this time, argon was mixed with nitrogen and
used chiefly for filling incandescent lamps; later, similar mixtures were used in fluor-
escent lamps. After 1943, demand for argon as a shield gas in arc welding grew to
exceed all other uses.

3.4.1. Properties of Argon


Table 3.32 gives the physical constants for argon. Argon is considered to be a very
inert gas and is not known to form true chemical compounds as do krypton, xenon,
and radon. Argon is useful chiefly for its inertness, particularly at high temperatures;
it is used to control reaction rates, exclude air from industrial processes and devices,
and mix and stir solutions or melts. No other gas is at once as effective and as eco-
nomical in these uses as argon. There is no problem of resource depletion because the

Table 3.27 The Heat of Vaporization of Oxygen–Nitrogen Mixtures at 101.3 kPa (1 atm)

Oxygen Heat Oxygen Heat Oxygen Heat


(mol T) (cal=g) (mol%) (cal=g) (mol%) (cal=g)

0 47.74 35 49.83 70 51.08


5 48.06 40 50.07 75 51.17
10 48.37 45 50.28 80 51.23
15 48.68 50 50.48 85 51.23
20 48.98 55 50.67 90 51.18
25 49.28 60 50.83 95 51.12
30 49.57 65 50.97 100 51.01

Note: There is a maximum at 82% oxygen. Argon is usually considered part of the nitrogen. The heat of a
21% oxygen mixture is 49.04 cal=g.

© 2005 by Marcel Dekker


5367-4 Flynn Ch03 R2 090804

Cryogenic Fluids 149

Table 3.28 Air Vapor–Liquid Equilibria—Selected Values

Sat. liquid (bubble) Sat. vapor (dew)


Pressure temperature temperature
 
atm psi K R K R

1 14.70 78.8 141.84 81.8 147.24


2 29.40 85.55 153.99 88.31 158.96
3 44.10 90.94 163.69 92.63 166.73
5 73.50 96.38 173.48 98.71 177.68
7 102.90 101.04 181.87 103.16 185.69
10 147.00 106.47 191.65 108.35 195.03
15 220.50 113.35 204.03 114.91 206.84
20 294.00 118.77 213.79 120.07 216.13
25 367.50 123.30 221.94 124.41 223.94
30 441.00 127.26 229.07 128.12 230.62
35 514.50 130.91 235.64 131.42 236.56
37.17 546.40 132.52 238.54 132.52 238.54
37.25 547.58 132.42 238.35 132.42 238.36

amount of argon available in the atmosphere is inexhaustible; it is not consumed in


use but returns to the atmosphere.
a. Heats of Vaporization and Fusion.

Units Triple point Boiling point

K 83.8 87.3
kJ=kg 27.78 161.6

b. Vapor Pressure. The vapor pressure of argon is given to a high degree of


accuracy (  0.02%) by the equation

log pðmmÞ ¼ A  B=T ð3:18Þ

where T is in kelvins. for liquid argon,

A ¼ 6:9224 ðfor 83:77  88:2 KÞ and B ¼ 352:8


For solid argon,

A ¼ 7:7353 ðfor 82  83:77 KÞ and B ¼ 420:9

These data are from Johnson (1960).


c. Dielectric Constant. The Clausius–Mosotti function
 
e1 1
ð3:19Þ
eþ2 r

should be a constant and independent of temperature provided the molecules of


the liquid studied have no permanent dipole moment. For liquid argon this quantity

© 2005 by Marcel Dekker


5367-4 Flynn Ch03 R2 090804

150 Chapter 3

Figure 3.56 Vapor pressure of air.

is constant within the accuracy of the data. e is the dielectric constant and r is the
density.
For liquid argon, values for the dielectric constant are given in Table 3.33
For gaseous argon at STP, according to Jelatis (1948),

e ¼ 1:000554

d. Density. The data tabulated below are from the work of Baly and Donnan
(1902). These results may be expressed by the formula

d ¼ 1:42333  0:006467ðT  84Þ

© 2005 by Marcel Dekker


5367-4 Flynn Ch03 R2 090804

Cryogenic Fluids 151

Table 3.29 Thermal Conductivity of Gaseous Air

Temperature Thermal conductivity



K R Watt=(cm K) Btu=(ft hr  R)

80 144 7.464  10–2 4.313  10–3


90 162 8.350  10–2 4.825  10–3
100 180 9.248  10–2 5.344  10–3
110 198 1.015  10–1 5.863  10–3
120 216 1.105  10–1 6.383  10–3
130 234 1.194  10–1 6.902  10–3
140 252 1.284  10–1 7.419  10–3
150 270 1.373  10–1 7.933  10–3
160 288 1.461  10–1 8.442  10–3
170 306 1.549  10–1 8.953  10–3
180 324 1.637  10–1 9.458  10–3
190 342 1.723  10–1 9.957  10–3
200 360 1.809  10–1 1.045  10–2
210 378 1.894  10–1 1.094  10–2
220 396 1.978  10–1 1.143  10–2
230 414 2.062  10–1 1.192  10–2
240 432 2.145  10–1 1.239  10–2
250 450 2.227  10–1 1.287  10–2
260 468 2.308  10–1 1.334  10–2
270 486 2.388  10–1 1.380  10–2
280 504 2.467  10–1 1.426  10–2
290 522 2.547  10–1 1.472  10–2
300 540 2.624  10–1 1.516  10–2

Source: Hilsenrath et al. (1955).

Table 3.30 Density of Saturated Liquid Air

Density
3
Pressure (atm) Temp. (K) Volume (cm =mol) g=cm3 lb=ft3

1 78.8 33.14 0.8739 54.56


2 85.55 34.39 0.8421 52.57
3 90.94 35.40 0.8181 51.07
5 96.38 36.94 0.7840 48.94
7 101.04 38.21 0.7579 47.31
10 106.47 40.00 0.7240 45.20
15 113.35 43.21 0.6702 41.84
20 118.77 46.63 0.6211 38.77
25 123.30 50.37 0.5749 35.89
30 127.26 55.69 0.5200 32.46
35 130.91 64.90 0.4462 27.86
37.17 132.52 90.52 0.3199 19.97
37.25 132.42 88.28 0.3280 20.48

© 2005 by Marcel Dekker


152
Table 3.31 Values of Distribution Coefficient Functions, In (Kp=po), for Nitrogen, Oxygen, and Argon. The Value of the Reference Pressure is
po ¼ 101.3 kPa (1 atm)

Nitrogen Oxygen Argon

Temp. (K) 101.3 kPa 202.6 kPa 506.6 kPa 101.3 kPa 202.6 kPa 506.6 kPa 101.3 kPa 202.6 kPa 506.6 kPa

78 0.0798 – – –1.3368 – – –0.9075 – –


80 0.3040 – – –1.1164 – – –0.7157 – –
82 0.5282 – – –0.8959 – – –0.5238 – –
84 0.7582 0.7048 – –0.6755 –0.4573 – –0.3319 –0.2516 –
86 0.9766 0.9030 – –0.4550 –0.3016 – –0.1400 –0.0717 –
88 1.2008 1.1012 – –0.2346 –0.1459 – –0.0519 þ0.1082 –
90 1.4249 1.2994 – –0.0141 þ0.0098 – þ0.1400 0.2880 –
92 – 1.4976 – – 0.1655 – – 0.4679 –
94 – 1.6958 1.5503 – 0.3211 0.6605 – 0.6477 0.5518
96 – 1.8939 1.7017 – 0.4768 0.7877 – 0.8276 0.7323
98 – – 1.8531 – – 0.9148 – – 0.8629
100 – – 2.0045 – – 1.0420 – – 1.0434
102 – – 2.1559 – – 1.1692 – – 1.2240
104 – – 2.3073 – – 1.2963 – – 1.4045
106 – – 2.4588 – – 1.4235 – – 1.5851
108 – – 2.6102 – – 1.5506 – – 1.7656

Chapter 3
© 2005 by Marcel Dekker
5367-4 Flynn Ch03 R2 090804

Cryogenic Fluids 153

where d is the density in g=cm3 and T is the absolute temperature in kelvins:

Temperature (K) Density (g=cm3)

84.0 1.4233
84.5 1.4201
85.0 1.4169
85.5 1.4136
86.0 1.4104
86.5 1.4072
87.0 1.4039
87.5 1.4007
88.0 1.3975
88.5 1.3942
89.0 1.3910
89.5 1.3878
90.0 1.3845

e. Other Properties. Figures 3.58–3.66 give the density, specific heats, melting
curve, surface tension, and viscosity of argon, taken from the WADD report men-
tioned before (Johnson, 1960)

3.4.2. Uses of Argon


Demand for argon has grown rapidly as new applications have developed. Gas-
shielded arc welding was once the major consumer. The principal use of argon
now is in the argon–oxygen decarburization (AOD) process in the stainless steel
industry. Argon is also used in metal working and in electrical and electronic applica-
tions. In 1983, US argon production was estimated to be 1.8 Mg (1800 metric tons)
per day. The US industry is currently operating at 100% of practical capacity,
primarily due to the demand of the AOD process.
The uses of argon result largely from its property of inertness in the presence of
reactive substances but also from its low thermal conductivity, low ionization poten-
tial, and good electrical conductivity.
a. Fabricated Metal Products. Argon is used in electric arc welding as an inert
gas shield, alone or mixed with other gases. It is used both with consumable electro-
des, most often similar in composition to the material being welded, or with noncon-
sumable tungsten electrodes in order to protect the weld from the atmosphere. It also
acts as a conductor to carry the arc welding current. It is especially suitable for
metals such as aluminum, stainless steel, titanium, and copper that are difficult to
weld.
Approximately 26% of the argon produced is used as a shielding gas in welding
fabricated metal products.
b. Iron and Steel Production. Argon is used to mix and degas carbon steel and
steel castings, to purge molds, and to provide neutral atmospheres in heat-treating
furnaces. As mentioned, the AOD process in the stainless steel industry is now the
largest single user of argon.
c. Nonferrous Metals Production. The major use of argon in nonferrous
metals production is to furnish an inert atmosphere in vessels in which titanium
tetrachloride is reduced to metal sponge by the Kroll process. Argon is also used

© 2005 by Marcel Dekker


5367-4 Flynn Ch03 R2 090804

154 Chapter 3

Figure 3.57 Temperature–entropy chart for argon.

in similar treatment of other metals, such as molybdenum, vanadium, and cesium,


that are reactive at the temperatures of preparation.
d. Chemicals. Argon is used for blanketing and atmosphere control in the
production of chemicals. It prevents reaction with elements in the air and, used as
a dilutent, reduces in a controlled fashion the rate of reactions involving gases.
e. Electronic Components. Argon atmospheres are used in integrated circuit
manufacturing, in single-crystal-growing furnaces, and as a carrier gas in construct-
ing the circuits.

© 2005 by Marcel Dekker


5367-4 Flynn Ch03 R2 090804

Cryogenic Fluids 155

Table 3.32 Physical Constants of Argon

Argon
Chemical Symbol: Ar
Synonym: LAR (liquid only)
CAS Registry Number: 7440–37–1
DOT Classification: Nonflammable gas
DOT Label: Nonflammable gas
Transport Canada Classification: 2.2
UN Number: UN 1006 (compressed gas); UN 1951 (refrigerated liquid)

US Units SI Units

Physical constants
Chemical formula Ar Ar
Molecular weight 39.95 39.95
Density of the gas at 70 F (21.1 C) 0.103 lb=ft3 1.650 kg=m3
and 1 atm
Specific gravity of the gas at 70 F 1.38 1.38
(21.1 C) and 1 atm
Specific volume of the gas at 70 F 9.71 ft3=lb 0.606 m3=kg
(21.1 C) and 1 atm
Density of the liquid at boiling point 87.02 lb=ft3 1394 kg=m3
and 1 atm
Boiling point at 1 atm 302.6 F 185.9 C
Melting point at 1 atm 308.6 F 189.2 C
Critical temperature 188.1 F 122.3 C
Critical pressure 711.5 psia 4905 kPa abs
Critical density 33.44 lb=ft3 535.6 kg=m3
Triple point 308.8 F at 9.99 psia 199.3 C at 68.9 kPa abs
Latent heat of vaporization at boiling 69.8 Btu=lb 162.3 kJ=kg
point and 1 atm
Latent heat of fusion at triple point 12.8 Btu=lb 29.6 kJ=kg
Specific heat of the gas at 70 F (21.1 C)
and 1 atm
CP 0.125 Btu=(lb)( F) 0.523 kJ=(kg)( C)
CV 0.075 Btu=(lb)( F) 0.314 kJ=(kg)( C)
Ratio of specific heats 1.67 1.67
Solubility in water at 32 F (0 C) vol=vol 0.056 0.056
Weight of the liquid at boiling point 11.63 lb=gal 1394.0 kg=m3

Table 3.33 Dielectric Constant for Argon—Selected Values

Temp. (K) Density (g=cm3) Dielectric constant (ref. to vacuum)

88.8 1.393 1.516


88.5 1.395 1.518
87.1 1.404 1.520
85.8 1.414 1.525
84.3 1.422 1.530
82.4a 1.434 1.537
a
Supercooled liquid.
Source: McLennan et al. (1930).

© 2005 by Marcel Dekker


5367-4 Flynn Ch03 R3 101104

156 Chapter 3

Figure 3.58 Density of solid argon.

f. Other Uses. About 1% of the demand for argon results from its use in
lamps. Infrared and photoflood lamps are filled with a mixture of about 88% argon
and 12% nitrogen. The high density and low reactivity of argon help retard the rate
of evaporation of the tungsten filament, thus prolonging the life of the lamp; the
nitrogen prevents arcing.

3.5. Neon
Neon (at wt. 20.183) has three stable isotopes of mass numbers 20, 21, and 22, which
occur in atmosphere air in the relative abundance 10,000:28:971. Five unstable iso-
topes are also known.
Neon is a rare gas element present in the atmosphere to the extent of 1 part in
65,000 of air. It is obtained by liquefaction of air and seperated from the other gases
by fractional distillation. It is a very insert element; however, it is said to form a com-
pound with fluorine. Of all the rare gases, the discharge of neon is the most intense at
ordinary voltages are currents.
Temperature entropy diagrams for neon are shown in Fig 3.67 and 3.68.

3.5.1. Properties of Neon


Liquid neon is a clear, colorless liquid that boils at 0.987 bar (1 atm) and 27.1 K
(48.8 R) and has a triple point at 24.56 K.
Table 3.34 gives the physical constants for neon, as well as krypton and xenon.
a. Phase Transition. See Table 3.35.
b. Heat of Vaporization and Heat of Fusion. See Table 3.36. The heat of fusion
of neon at the melting point is 76.67 kJ=kg (3.969 cal=g; 80.1 cal=mol).

© 2005 by Marcel Dekker


5367-4 Flynn Ch03 R2 090804

Cryogenic Fluids 157

Figure 3.59 Density of saturated liquid argon.

© 2005 by Marcel Dekker


5367-4 Flynn Ch03 R2 090804

158 Chapter 3

Figure 3.60 Density of gaseous argon.

© 2005 by Marcel Dekker


5367-4 Flynn Ch03 R2 090804

Cryogenic Fluids 159

Figure 3.61 Specific heat of liquid argon.

Figure 3.62 Specific heat of solid argon

© 2005 by Marcel Dekker


5367-4 Flynn Ch03 R2 090804

160 Chapter 3

Figure 3.63 Melting curve for argon.

© 2005 by Marcel Dekker


5367-4 Flynn Ch03 R2 090804

Cryogenic Fluids 161

Figure 3.64 Surface tension of saturated liquid argon.

Figure 3.65 Viscosity of gaseous argon at 25 C

© 2005 by Marcel Dekker


5367-4 Flynn Ch03 R2 090804

162 Chapter 3

Figure 3.66 Viscosity of saturated liquid argon.

c. Melting Point Temperature of Neon. Mills and Grilley (1955) give the melt-
ing curve of neon as
p ¼ a þ bT c ð3:20Þ
where p is the pressure in kg=cm2, T is the temperature in K, a ¼ 1057.99,
b ¼ 6.289415, c ¼ 1.599916.
d. Vapor Pressure of Solid Neon. Between 15 and 20.5 K, the following
equation may be used:
111:76
log np ¼  þ 6:0424 ð3:21Þ
T
where vp is in cmHg and T is in K.
e. Vapor Pressure of Liquid Neon. Verschaffelt (1929) gives the following
vapor pressure equation for liquid neon between 25 and 35 K:
T log np ¼28:100 þ 36:00ðT  35Þ  0:003333ðT  35Þ2
ð3:22Þ
þ 0:000400ðT  35Þ3 þ 0:00002667ðT  35Þ4
where vp is in atm and T is in K on the Leiden scale (0 C ¼ 273.09 K).
f. Density. The density of solid neon (Johnson, 1960) is in the range:

Temp. (K) Density (kg=m3)

4.3 1443
24.57a 1444
a
Normal melting point value.

© 2005 by Marcel Dekker


5367-4 Flynn Ch03 R2 090804

Cryogenic Fluids 163

Figure 3.67 Temperature–entropy chart for neon, 25–80 K.

© 2005 by Marcel Dekker


5367-4 Flynn Ch03 R2 090804

164 Chapter 3

Figure 3.68 Temperature–entropy chart for neon, 60–300 K.

© 2005 by Marcel Dekker


5367-4 Flynn Ch03 R2 090804

Cryogenic Fluids 165

Table 3.34 Physical Constants of Krypton, Neon, and Xenon

Rare Gases: Krypton, Neon, Xenon


Krypton Neon Xenon

Chemical Symbol: Kr Ne Xe
CAS Registry Number: 7439-90-9 7440-01-9 7440-63-3
DOT Classification: Nonflammable gas Nonflammable gas Nonflammable gas
DOT Label: Nonflammable gas Nonflammable gas Nonflammable gas
Transport Canada 2.2 2.2 2.2
Classification:
UN Number:
(compressed gas): UN 1056 UN 1065 UN 2036
(refrigerated liquid): UN 1970 UN 1913 UN 2591

US Units SI Units

Krypton
Chemical formula Kr Kr
Molecular weight 83.80 83.80
Density of the gas at 70 F 0.2172 lb=ft3 3.479 kg=m3
(21.1 C) and 1 atm
Specific gravity of the gas at 2.899 2.899
70 F (21.1 C) and 1 atm
(air ¼ 1)
Specific volume of the gas at 4.604 ft3=lb 0.287 m3=kg
70 F (21.1 C) and 1 atm
Boiling point at 1 atm 244.0 F 153.4 C
Melting point at 1 atm 251 F 157 C
Critical temperature 82.8 F 63.8 C
Critical pressure 798.0 psia 5502 kPa abs
Critical density 56.7 lb=ft3 908 kg=m3
Triple point 251.3 F at 10.6 psia 157.4 C at
73.2 kPa abs
Latent heat of vaporization 46.2 Btu=lb 107.5 kJ=kg
at boiling point
Latent heat of fusion at 8.41 Btu=lb 19.57 kJ=kg
triple point
Specific heat of the gas at 70 F
(21.1 C) and 1 atm
CP 0.060 Btu=(lb)( F) 0.251 kJ=(kg)( C)
CV 0.035 Btu=(lb)( F) 0.146 kJ=(kg)( C)
Ratio of specific heats, CP=CV 1.69 1.69
Solubility in water, vol=vol at 0.0594 0.0594
68 F (20 C)
Weight of the liquid at boiling point 20.15 lb=gal 2415 kg=m3
Density of the liquid at boiling point 150.6 lb=ft3 2412.38 kg=m3
Gas=liquid volume ratio (liquid 693.4 693.4
at boiling point, gas at 70 F
and 1 atm, vol=vol)

Neon
Physical Constants
Chemical formula Ne Ne
Molecular weight 20.183 20.183
Density of the gas at 70 F (21.1 C) 0.05215 lb=ft3 0.83536 kg=m3
and 1 atm
Specific gravity of the gas at 70 F (21.1 C) 0.696 0.696
and 1 atm (air ¼ 1)

(Continued)

© 2005 by Marcel Dekker


5367-4 Flynn Ch03 R2 090804

166 Chapter 3

Table 3.34 (Continued )

US Units SI Units

Specific volume of the gas at 70 F (21.1 C) 19.18 ft3=lb 1.197 m3=kg
and 1 atm
Boiling point at 1 atm 410.9 F 246.0 C
Melting point at 1 atm 415.6 F 248.7 C
Critical temperature 379.8 F 228.8 C
Critical pressure 384.9 psia 2654 kPa abs
Critical density 30.15 lb=ft3 483 kg=m3
Triple point 415.4 F at 6.29 psia 248.6 C at 43.4 kPa abs
Latent heat of vaporization at boiling point 37.08 Btu=lb 86.3 kJ=kg
Latent heat of fusion at triple point 7.14 Btu=lb 16.6 kJ=kg
Specific heat of the gas at 70 F (21.1 C)
and 1 atm
CP 0.25 Btu=(lb)( F) 1.05 kJ=(kg)( C)
CV 0.152 Btu=(lb)( F) 0.636 kJ=(kg)( C)
Ratio of specific heats, CP=CV 1.64 1.64
Solubility in water, vol=vol at 68 F (20 C) 0.0105 0.0105
Weight of the liquid at boiling point 10.07 lb=gal 1207 kg=m3
Density of saturated vapor at 1 atm 0.5862 lb=ft3 9.390 kg=m3
Density of the gas at boiling point 0.6068 lb=ft3 9.7200 kg=m3
Density of the liquid at boiling point 75.35 lb=ft3 1207 kg=m3
Gas=liquid volume ratio (liquid at boiling point, 1445 1445
gas at 70 F and 1 atm, vol=vol)

Xenon
Physical Constants
Chemical formula Xe Xe
Molecular weight 131.3 131.3
Density of the gas at 70 F (21.1 C) and 1 atm 0.3416 lb=ft3 5.472 kg=m3
Specific gravity of the gas at 70 F and 4.560 4.560
1 atm (air ¼ 1)
Specific volume of the gas at 70 F (21.1 C) 2.927 ft3=lb 0.183 m3=kg
and 1 atm
Boiling point at 1 atm 162.6 F 108.2 C
Melting point at 1 atm 168 F 111 C
Critical temperature 61.9 F 16.6 C
Critical pressure 847.0 psia 5840 kPa abs
Critical density 68.67 lb=ft3 1100 kg=m3
Triple point 169.2 F at 11.84 psia 111.8 C at 81.6 kPa abs
Latent heat of vaporization at boiling point 41.4 Btu=lb 96.3 kJ=kg
Latent heat of fusion at triple point 7.57 Btu=lb 17.6 kJ=kg
Specific heat of the gas at 70 F (21.1 C)
and 1 atm
CP 0.038 Btu=(lb)( F) 0.269 kJ=(kg)( C)
CV 0.023 Btu=(lb)( F) 0.096 kJ=(kg)( C)
Ratio of specific heats, CP=CV 1.667 1.667
Solubility in water, vol=vol at 68 F (20 C) 0.108 0.108
Weight of the liquid at boiling point 25.51 lb=gal 3057 kg=m3
Density of the liquid at boiling point 190.8 lb=ft3 3057 kg=m3
Gas=liquid volume ratio (liquid at 558.5 558.5
boiling point, gas at 70 F and 1 atm,
vol=vol)

© 2005 by Marcel Dekker


5367-4 Flynn Ch03 R2 090804

Cryogenic Fluids 167

Table 3.35 Phase Transition Temperatures of Neon

Pressure Temperature

Property kPa atm psia K R

Critical pointa 2719 26.84 394.5 44.38 79.88


Boiling point 101.3 1 14.696 27.17 48.91
Melting point 101.3 1 14.696 24.57 44.23
Triple point 43.1 0.4257 6.256 24.57 44.23
a
Critical volume ¼ 42 cm3=mol.

Table 3.36 Latent Heat of Vaporization of Neon

Temperature Heat of vaporization



K R cal=g kJ=kg

25.17 45.31 21.36 89.37


26.15 47.07 20.96 87.70
27.15 48.87 20.56 86.02
27.17 48.91 20.6 86.19
30.13 54.23 19.34 80.92
33.09 59.56 17.97 75.19
36.05 64.89 16.23 67.91
37.83 68.09 14.87 62.22
39.08 70.34 13.69 57.28
41.065 73.92 11.26 47.11
43.02 77.44 7.491 31.34

The density of liquid neon is tabulated below.

Temp. (K) Density (kg=m3)

24.57a 1248
25.17 1238
27.17b 1204
30.13 1149
36.05 1017
39.08 928
43.02 749
44.38c 483
a
Triple point temperature.
b
Normal boiling temperature.
c
Critical temperature.

Symbols: rsat ¼ saturation pressure or vapor pressure, r ¼ density, CzP ¼ speci-


specific heat at constant pressure, m ¼ viscosity, k ¼ thermal conductivity, hfg ¼ heat
of vaporization, Npr ¼ mCp=K ¼ Prandtl number, sL ¼ surface tension.

© 2005 by Marcel Dekker


5367-4 Flynn Ch03 R2 090804

168 Chapter 3

g. Dielectric Constant of Gaseous Neon. One value has been published at 70 F
and 1 atm. The Clausius–Mosotti function may be used to estimate values at other
conditions with e ¼ 1.0001274.
h. Other Properties. Figures 3.69–3.76 give the density, specific heats, thermal
conductivity, vapor pressure, and melting curve of neon, taken from the WADD
report mentioned before (Johnson 1960).

3.5.2. Uses of Neon


Neon is used in making the common neon advertising signs, which accounts for its
largest use. It is also used to make high-voltage indicators, lightning arrestors, wave-
meter tubes, and TV tubes.
Liquid neon is now commercially available and is finding important applica-
tion as an economical cryogenic refrigerant. It has over 40 times more refrigerating
capacity per unit volume than liquid helium and more than three times that of liquid
hydrogen. It is compact, inert, and less expensive than helium when it meets refrig-
eration requirements.

Figure 3.69 Density of saturated liquid neon.

© 2005 by Marcel Dekker


5367-4 Flynn Ch03 R2 090804

Cryogenic Fluids 169

Figure 3.70 Density of saturated vapor neon.

© 2005 by Marcel Dekker


5367-4 Flynn Ch03 R2 090804

170 Chapter 3

Figure 3.71 Specific heat of saturated liquid neon.

Figure 3.72 Specific heat of solid neon.

© 2005 by Marcel Dekker


5367-4 Flynn Ch03 R2 090804

Cryogenic Fluids 171

Figure 3.73 Thermal conductivity of gaseous neon at 1 atm.

© 2005 by Marcel Dekker


5367-4 Flynn Ch03 R2 090804

172 Chapter 3

Figure 3.74 Vapor pressure of liquid neon.

© 2005 by Marcel Dekker


5367-4 Flynn Ch03 R2 090804

Cryogenic Fluids 173

Figure 3.75 Vapor pressure of solid neon.

© 2005 by Marcel Dekker


5367-4 Flynn Ch03 R2 090804

174 Chapter 3

Figure 3.76 Melting curve for neon.

© 2005 by Marcel Dekker


5367-4 Flynn Ch03 R2 090804

Cryogenic Fluids 175

3.6. Fluorine
Liquid fluorine (at wt. 18.9984) is a light yellow liquid having a normal boiling point
of 85.3 K (153.6 R). At 53.5 K (96.5 R) and 0.987 bar (1 atm), liquid fluorine freezes
as a yellow solid, but upon subcooling to 45.6 K (82 R), it transforms to a white
solid. Liquid fluorine is one of the most dense cryogenic liquids [density at normal
boiling point ¼ 1504.3 kg=m3 (93.9 lbm=ft3)].
Fluorine will react with almost all substances. If it comes in contact with
hydrocarbons, it will react hypergolically with a high heat of reaction, which is some-
times sufficiently high that the metal container for the fluorine is ignited. Satisfactory
handling systems for liquid fluorine have been operated, but this substance presents
even greater difficulty than oxygen because of its higher reactivity. The flow of liquid
fluorine through a system is preceded by a process of ‘‘passivation’’ in which fluorine
gas is admitted slowly to the system in order to build up a layer of passive fluorides
on its surfaces. This layer then presents a chemically inert barrier to the large quan-
tities of fluorine that follow.
Fluorine is highly toxic. The fatal concentration range for animals is
200 ppm hr; i.e., for an exposure time of l hr, 200 ppm of fluorine is fatal; for an expo-
sure time of 15 min, 800 ppm is fatal; and for an exposure time of 4 hr, 50 ppm is
fatal. The maximum allowable concentration for human exposure is usually consid-
ered to be approximately 1 ppm hr. The presence of fluorine in air may be detected by
its sharp, pungent odor for concentrations as low as 1–3 ppm.
Fluorine was first isolated in 1886 by the French chemist Henri Moissan after
nearly 75 years of unsuccessful attempts by several others. For many years after its
isolation, fluorine remained little more than a scientific curiosity, to be handled with
extreme caution because of its toxicity. Commercial production of fluorine began
during World War II when large quantities were required in the fluorination of ura-
nium tetrafluoride (UF4) to produce uranium hexafluoride (UF6) for the isotopic
separation of uranium-235 by gaseous diffusion in the development of the atomic
bomb. Today, commercial production methods are essentially variations of the
Moisson process, and safe techniques have been developed for the bulk handling
of liquid fluorine.
The major use of fluorine is in the form of fluorspar. Although not a major
commodity in terms of total quantity produced, fluorspar is a critical raw material
for the aluminum, chemical, and steel industries of the world. There is at present
no significant use of cryogenic liquid fluorine per se.
Table 3.37 gives the physical constants for fluorine.

3.6.1. Phase Transition


The phase transition heats of fluorine are listed in Table 3.38.

3.6.2. Heat of Vaporization


See Table 3.39

3.6.3. Vapor Pressure


Measurements of the vapor pressure have been made by many authors; agreement is
good. Hu et al. (1953), whose data extend over a wide range give
357:258 1:3155  1013
log P ¼ 7:08718   ð3:23Þ
T T3
where P is in mmHg and T is in K.

© 2005 by Marcel Dekker


5367-4 Flynn Ch03 R2 090804

176 Chapter 3

Table 3.37 Physical Constants of Fluorine

Chemical Symbol: F2
CAS Registry Number: 7782-41-4
DOT Classification: Nonflammable gas
DOT Label: Poison and Oxidizer
Transport Canada Classification: 2.3 (5.1)
UN Number: UN 1045

U.S.Units SI Units

Physical constants
Chemical formula F2 F2
Molecular weight 38.00 38.00
Density of the gas
at 32 F (0.0 C) and 1 atm 0.106 lb=ft3 1.70 kg=m3
at 70 F (21.1 C) and 1 atm 0.098 lb=ft3 1.57 kg=m3
Specific gravity of the gas
at 70 F (21.1 C) and 1 atm 1.312 1.312
Specific volume of the gas
at 70 F (21.1 C) and 1 atm 10.17 0.635
Density of the liquid
at 306.8 F (188.2 C) 94.2 lb=ft3 1509 kg=m3
at 320.4 F (195.8 C) 97.9 lb=ft3 1568 kg=m3
Boiling point at 1 atm 306.8 F 188.2 C
Melting point at 1 atm 363.4 F 219.6 C
Critical temperature 199.9 F 128.8 C
Critical pressure 756.4 psia 5215 kPa abs
Critical density 35.8 lb=ft3 573.6 kg=m3
Triple point 363.4 F at 219.6 C at
0.324 psia 0.223 kPa abs
Latent heat of vaporization
at 306.8 F (188.2 C) 74.8 Btu=lb 173.5 kJ=kg
Latent heat of fusion
at 363.4 F (219.6 C) 5.8 Btu=lb 13.4 kJ=kg
Weight of liquid per gallon
at 306.8 F (188.1 C) 12.6 lb=gal 1509.8 kg=m3
at 320.4 F (195.8 C) 13.1 lb=gal 1569.7 kg=m3
Heat capacity of the gas, CP
at 32 F (0 C) 0.198 Btu=(lb)( F) 0.828 kJ=(kg)( C)
at 70 F (21.1 C) 0.197 Btu=(lb)( F) 0.825 kJ=(kg)( C)
Heat capacity of the gas, CV
at 70 F (21.1 C) 0.146 Btu=(lb)( F) 0.610 kJ=(kg)( C)
Ratio of specific heats, CP=CV
at 70 F (21.1 C) 1.353 1.353
Thermal conductivity of gas
at 32 F (0 C) and 1 atm 0.172 Btu.in.=hr F 0.248 W=m.K
Viscosity
of gas at 32 F (0 C) and 1 atm 0.0218 cP 0.0218 mPa s
of liquid at 314 F (192.2 C) 0.275 cP 0.275 mPa s

© 2005 by Marcel Dekker


5367-4 Flynn Ch03 R2 090804

Cryogenic Fluids 177

Table 3.38 Phase Transition Heats of Fluorine

Transition Temp. (K) Transition heats (kJ=kg)

Solid transition 45.55 19.14


Fusion 53.51 13.43
Vaporization 85.21 166.22

Source: Hu et al. (1953); Rossini et al. (1952).

Table 3.39 Heat of Vaporization of Fluorine

Temp. (K) Heat of vaporization (kJ=kg)

85.25 166.2
90 160.9
95 154.9
100 148.6
105 142.0
110 134.6
115 126.6
120 117.7
125 107.6
130 95.65
135 80.81
140 59.12
144 14.85

Source: Fricke (1948).

3.6.4. Density of Liquid Fluorine


The data of Jarry and Miller (1956) are represented by the following equation and
are tabulated below.
r ¼ 1:907  2:201  103 T  2:948  105 T 2 ð3:24Þ
where r (density) is in g=cm3 and T is in K.

Density of Fluorine—Selected Values

Temp. (K) Density (kg=m3) Temp. (K) Density (kg=m3)

65.78 1638 86.91 1496


71.76 1594 88.26 1481
74.93 1578 88.50 1484
78.59 1550 90.08 1472
78.62 1553 91.55 1458
81.72 1532 91.75 1460
81.73 1528 94.73 1434
84.34 1514 97.56 1412
85.05 1505 100.21 1391
85.67 1505 102.75 1370
Source: Jarry and Miller (1956).

© 2005 by Marcel Dekker


5367-4 Flynn Ch03 R2 090804

178 Chapter 3

Table 3.40 Viscosity of Gaseous Fluorine

Temp. (K) Viscosity (cP)

90.0 0.767
169.3 1.424
200.0 1.680
289.1 2.345
327.1 2.547

Source: Franck and Stober (1952).

3.6.5. Viscosity
Tables 3.40 and 3.41 present data on the viscosity of gaseous and liquid fluorine,
respectively.

3.6.6. Thermal Conductivity


For data on the thermal conductivity of gaseous fluorine, see Table 3.42.
Symbols: rsat ¼ saturation pressure or vapor pressure, r ¼ density, CP ¼ specific
heat at constant pressure, m ¼ viscosity, k ¼ thermal conductivity, nfg ¼ heat of
vaporization, Npr ¼ mCP=K ¼ Prandtl number, sL ¼ surface tension.

3.6.7. Other Properties


Figures 3.77–3.82 give the specific heats, density, heat of vaporization, vapor pres-
sure, surface tension, and viscosity of fluorine, taken from the WADD report men-
tioned before (Johnson, 1960).

Table 3.41 Viscosity of Liquid Fluorine

Temp. (K) Viscosity (cP) Temp. (K) Viscosity (cP)

69.2 0.414 78.2 0.299


73.2 0.349 80.9 0.275
75.3 0.328 83.2 0.257

Source: Elverum and Doescher (1952).

Table 3.42 Thermal Conductivity of Gaseous Fluorine

Temperature (K) Thermal conductivity (mW=(cm K))

100 0.0862
150 0.134
200 0.183
250 0.228
273 0.247
300 0.269
350 0.308

Source: Franck and Wicke (1951).

© 2005 by Marcel Dekker


5367-4 Flynn Ch03 R2 090804

Cryogenic Fluids 179

Figure 3.77 Specific heat (CP) of gaseous fluorine at 1 atm.

© 2005 by Marcel Dekker


5367-4 Flynn Ch03 R2 090804

180 Chapter 3

Figure 3.78 Density of liquid fluorine.

3.7. Hydrogen
Natural hydrogen (at wt. 1.008) is a mixture of two stable isotopes; hydrogen (1H),
of atomic mass 1, and deuterium (2H), of atomic mass 2. The abundance ratio is
about 6400:1, although slight variations occur, depending on the source of the
hydrogen. Actually, since molecular hydrogen is diatomic, nearly all the deuterium
atoms in natural hydrogen exist in combination with hydrogen atoms. Accordingly,
ordinary hydrogen is a mixture of H2 and HD (1H2H) molecules in the ratio 3200:1.
A very rare radioactive isotope of hydrogen of atomic mass 3 (3H), called tritium,
also exists. Tritium has a half-life of about 12.5 years.
Hydrogen is the most abundant of all the elements in the universe, and it is
thought that the heavier elements were, and still are, being built from hydrogen
and helium. It has been estimated that hydrogen makes up more than 90% of all
the atoms or three-fourths of the mass of the universe. Production of hydrogen in
the United States alone now amounts to 85 million m3 (3 billion ft3). The physical
constants for hydrogen are shown in Table 3.43.

3.7.1. Large-Scale Liquid Hydrogen Production


From the time hydrogen was first liquefied by James Dewar in 1898 until the 1940s,
liquid hydrogen remained a laboratory curiosity, with production rates measured in a
few grams. During the mid-1940s to mid-1950s, substantial-scale laboratory hydro-
gen liquefiers were built to meet the needs of the US government’s nuclear and aero-
space programs, with the largest such liquefier being a 240 L=hr (0.41 metric ton=day)
unit at the National Bureau of Standards in Boulder, Colorado. The first commercial-
scale liquid hydrogen production plants were built in the late 1950s under a then
secret US. Air Force program. Under contract to the Air Force, three successively
larger liquid hydrogen plants code named Baby Bear (0.75 US tons=day or 0.68

© 2005 by Marcel Dekker


5367-4 Flynn Ch03 R2 090804

Cryogenic Fluids 181

Figure 3.79 Heat of vaporization of fluorine.

© 2005 by Marcel Dekker


5367-4 Flynn Ch03 R2 090804

182 Chapter 3

Figure 3.80 Vapor pressure of liquid fluorine.

© 2005 by Marcel Dekker


5367-4 Flynn Ch03 R2 090804

Cryogenic Fluids 183

Figure 3.81 Surface tension of liquid fluorine.

© 2005 by Marcel Dekker


5367-4 Flynn Ch03 R2 090804

184 Chapter 3

Figure 3.82 Viscosity of gaseous fluorine.

© 2005 by Marcel Dekker


5367-4 Flynn Ch03 R2 090804

Cryogenic Fluids 185

Table 3.43 Physical Constants of Hydrogen

Hydrogen
Chemical Symbol: H2
CAS Registry Number: 1333-74-0
DOT Classification: Flammable gas
DOT Label: Flammable gas
Transport Canada Classification: 2.1
UN Number: UN 1049 (compressed gas); UN 1966 (refrigerated liquid)

U.S. Units SI Units

Normal Hydrogen
Chemical formula H2 H2
Molecular weight 2.016 2.016
Density of the gas at 70 F (21.1 C) 0.00521 lb=ft3 0.08342 kg=m3
and 1 atm
Specific gravity of the gas at 32 F 0.06960 0.06960
(0 C) and 1 atm (air ¼ 1)
Specific volume of the gas at 70 F 192.0 ft3=lb 11.99 m3=kg
(21.1 C) and 1 atm
Boiling point at 1 atm 423.0 F 252.8 C
Melting point at 1 atm 434.55 F 259.2 C
Critical temperature 399.93 F 239.96 C
Critical pressure 190.8 psia 1315 kPa abs
Critical density 1.88 lb=ft3 30.12 kg=m3
Triple point 434.55 F at 259.2 C at
1.045 psia 7.205 kPa abs
Latent heat of vaporization at 191.7 Btu=lb 446.0 kJ=kg
boiling point
Latent heat of fusion at triple point 24.97 Btu=lb 58.09 kJ=kg
Specific heat of the gas at 70 F
(21.1 C) and 1 atm
CP 3.425 Btu=(lb)( F) 14.34 kJ=(kg)( C)
CV 2.418 Btu=(lb)( F) 10.12 kJ=(kg)( C)
Ratio of specific heats
(CP=CV) 1.42 1.42
Solubility in water, vol=vol at 60 F 0.019 0.019
(15.6 C)
Density of the gas at boiling point 0.083 lb=ft3 1.331 kg=m3
and 1 atm
Density of the liquid at boiling point 4.43 lb=ft3 (0.5922 lb=gal) 70.96 kg=m3
and 1 atm
Gas=liquid ratio (liquid at boiling point, 850.3 850.3
gas at 70 F (21.1 C) and 1 atm), vol=vol
Heat of combustion at 70 F (21.1 C)
and 1 atm
Gross 318.1 Btu=ft3 11,852 kJ=m3
Net 268.6 Btu=ft3 10,009 kJ=m3

Para Hydrogen
(The following constants are different from normal hydrogen)
Boiling point at 1 atm 423.2 F 252.9 C
Melting point 434.8 F 259.3 C

(Continued)

© 2005 by Marcel Dekker


5367-4 Flynn Ch03 R2 090804

186 Chapter 3

Table 3.43 (Continued )

Critical temperature 400.31 F 240.17 C


Critical pressure 187.5 psia 1293 kPa abs
Critical density 1.96 lb=ft3 31.43 kg=m3
Triple point 434.8 F at 259.3 C at
1.021 psia 7.042 kPa abs
Latent heat of vaporization at boiling point 191.6 Btu=lb 445.6 kJ=kg
Latent heat of fusion at triple point 25.06 Btu=lb 58.29 kJ=kg
Specific heat of the gas at 70 F (21.1 C)
and 1 atm
CP 3.555 Btu=(lb)( F) 14.88 kJ=(kg)( C)
CV 2.570 Btu=(lb)( F) 10.76 kJ=(kg)( C)
Ratio of specific heats
(CP=CV) 1.38 1.38
Density of the gas at boiling point and 1 atm 0.084 lb=ft3 1.338 kg=m3
Density of the liquid at boiling point and 1 atm 4.42 lb=ft3 70.78 kg=m3
(0.5907 lb=gal)
Gas=liquid ratio (liquid at boiling point, 848.3 848.3
gas at 70 F (21.1 C) and 1 atm, vol=vol

metric tons=day), Mama Bear (3.5 US tons=day or 3.2 metric tons=day), and Papa
Bear (30 tons=day or 27 metric tons=day) were designed and built by Air Products
and Chemicals, Inc. Papa Bear, which went on-stream in West Palm Beach, Florida,
in February 1959, was the world’s first truly large-scale liquid hydrogen plant.
From 1960 to the present, several additional large-scale commercial liquid
hydrogen plants were built in North America. For example, in 1964 the Linde Divi-
sion of Union Carbide built the largest single plant to date, a 60 US ton=day (54
metric tons=day) plant in Sacramento, California. Between 1965 and 1976, Air Pro-
ducts constructed a 66 US tons=day (60 metric tons=day) two-plant liquid hydrogen
production complex in New Orleans, Louisiana, and between 1982 and 1990 three
liquid hydrogen plants were built in Canada by Air Products, Liquid Air, and BOC.
Today, nine large-scale liquid hydrogen plants are in operation in the United
States and Canada with a combined capacity of about 190 US tons=day (172 metric
tons=day). Some older plants like Papa Bear and the Linde Sacramento plant no
longer exist. Table 3.44 lists some key milestones in the creation of large-scale com-
mercial liquid hydrogen plants, and Fig. 3.83 shows the location, size, and ownership
of liquid hydrogen plants in North America in 1990. The aerospace program, which
created the initial demand for large quantities of liquid hydrogen, accounted for only
20% of liquid hydrogen demand in 1990. A variety of commercial applications
account for the remaining demand.
The diversity of commercial liquid hydrogen containers for both stationary sto-
rage and transportation are listed below. Three insulation technologies are
employed. In increasing order of thermal efficiency and cost, they are evacuated per-
lite insulation, evacuated multilayer insulation, and liquid nitrogen shielded multi-
layer insulation. Large stationary storage containers are reasonably efficient, using
just an evacuated annulus filled with perlite, due to their relatively low surface-to-
volume ratio. Smaller stationary containers and all transportation containers require
multilayer insulation for acceptable efficiency (i.e., low vent rates). Liquid nitrogen
shielded multilayer insulation is widely used for liquid helium storage and transport,

© 2005 by Marcel Dekker


5367-4 Flynn Ch03 R2 090804

Cryogenic Fluids 187

Table 3.44 Milestones in the Development of Large-Scale Commercial Liquid Hydrogen


Plantsa

Date Event

1898 Hydrogen first liquefied by James Dewar


1898–mid-40s Laboratory curiosity
Mid 1940s–1956 Laboratory and pilot-scale liquefiers built to support
US nuclear weapons and aerospace programs. Largest unit:
0.45 ton=day at National Bureau of Standards
1956–1959 US Air Force and Air Products—‘‘Bear’’ program:
1957 Baby Bear—Painsville, OH (0.75 ton=day)
1957 Mama Bear—West Palm Beach, FL (3.5 tons=day)
1959 Papa Bear—West Palm Beach, FL (30 tons=day)
1960–present Several large-scale plants constructed in North America
through private funding and operation. Examples:
1964 Linde 60 tons=day LH2 plant in Sacramento, CA; largest ever built
1965, 1976 Air Products builds two-plant, 66 tons=day LH2 complex in
New Orleans, LA
1982–1990 Three LH2 plants built in Canada—Air Products, Liquid Air, Airco
1980s–present LH2 plants built in Europe and Japan
a
All plant capacities given in US tons=day.

but for liquid hydrogen it is typically limited to portable containers that must be
shipped long distances (often overseas) without venting for several days.
Boiloff from product liquid hydrogen storage tanks at the liquid hydrogen
plant site is usually redirected back to the liquefier plant for reliquefaction.

3.7.2. Molecular Forms of Hydrogen and Ortho–Para Conversion


One of the properties of hydrogen that sets it apart from other substances is the fact
that it can exist in two different molecular forms: orthohydrogen (o-H2) and para-
hydrogen (p-H2).

Liquid Hydrogen Transportation and Storagea

Container Capacity (US gal)b

Transportation
Over-the-road tankers 13,000–15,000
Seagoing containers 11,000
Inland waterway barges 250,000
Railcars 28,000–34,000
Storage
Large tanks
Spheres at Kennedy Space Center 830,000
Spheres at Air Products’ New Orleans plant 500,000
Horizontal cylinders at Air Products’
Sacramento plant 70,000
Customer tanks
Horizontal=vertical cylinder 1500–9000
Horizontal cylinder 9000–20,000
a
Insulation types include perlite, multilayer, and liquid nitrogen shielded multilayer.
b
1000 US gal ¼ 3.785 L.

© 2005 by Marcel Dekker


5367-4 Flynn Ch03 R2 090804

188 Chapter 3

Figure 3.83 Liquid hydrogen plants in the United States and Canada, 1990.

Figure 3.84 illustrates the two states of the hydrogen molecule: the higher
energy ortho state in which the protons in the two nuclei spin in the same direction,
and the lower energy para state in which the protons spin in opposite directions.
Temperature determines the relative abundance of each state, as shown in Table 3.45
and by the ortho–para thermodynamic equilibrium diagram in Fig. 3.85. At ambient
and higher temperatures, hydrogen consists of 25% para and 75% ortho and is
referred to as ‘‘normal’’ hydrogen (n-H2). At the normal boiling point of liquid
hydrogen (20.4 K), equilibrium shifts to almost 100% para. Conversion from ortho
to para is exothermic, with the heat of conversion of 302 Btu=lb (702 J=g) substan-
tially exceeding the latent heat of vaporization of normal boiling point hydrogen
of 195 Btu=lb (454 J=g).
The equilibrium mixture of o-H2 and p-H2 at any given temperature is called
equilibrium hydrogen (e-H2). At the normal boiling point of hydrogen, 20.4 K
(36.7 R), equilibrium hydrogen has a composition of 0.21% o-H2 and 99.79%
p-H2, or practically all parahydrogen.
Two deuterium atoms can combine to form a deuterium molecule, often
denoted D2 and sometimes called heavy hydrogen. As in the case of ‘‘light’’ hydro-
gen, the nuclear spins of the atoms in a deuterium molecule can assume different

Figure 3.84 Ortho- and parahydrogen differ in the manner in which their two protons spin
in the nucleus of the hydrogen atom.

© 2005 by Marcel Dekker


5367-4 Flynn Ch03 R2 090804

Cryogenic Fluids 189

Table 3.45 Ortho–Para Hydrogen Data

H2, % in D2, % in Heat of conversiona


Temperature (K) para form para form (cal=(g mol)

10 99.9999 0.0277 338.648


20 99.821 1.998 338.649
20.39 99.789 – 338.648
23.57 – 3.761 –
30 97.021 7.864 338.648
33.1 95.034 – 338.648
40 88.727 14.784 338.634
50 77.054 20.718 338.460
60 65.569 25.131 337.616
70 55.991 28.162 335.200
80 48.537 30.141 330.164
90 42.882 31.395 321.700
100 38.620 32.164 309.440
120 32.99 32.916 274.475
150 28.603 33.246 207.175
200 25.974 33.327 105.20
250 25.264 – 45.31
298.16 25.075 33.333 18.35
300 25.072 33.333 17.71
350 25.019 – –
400 25.005 – –
500 25.000 – –
a
For conversion of o-H2 to p-H2.

spatial orientations with respect to each other, although the orientations are a bit
more complicated. Ortho- and para deuterium have different sets of possible spatial
orientations, and the ortho–para equilibrium concentrations are temperature-
dependent, as in the case of hydrogen. The nucleus of the deuterium atom consists

Figure 3.85 Equilibrium para content of hydrogen vs. temperature.

© 2005 by Marcel Dekker


5367-4 Flynn Ch03 R2 090804

190 Chapter 3

of one proton and one neutron, so that the high-temperature composition (compo-
sition of normal deuterium) is two-thirds orthodeuterium and one-third paradeuter-
ium. In the case of deuterium, p-D2 converts to o-D2 as the temperature is
decreased, in contrast to hydrogen, in which o-H2 converts to p-H2 upon a decrease
in temperature. (The prefix ortho- simply means common and does not inherently
refer to the more energetic form.) The hydrogen deuteride molecule does not have
the symmetry that hydrogen and deuterium molecules possess; therefore, HD exists
in only one form.
Temperature–entropy diagrams for parahydrogen and deuterium are presented
in Figs. 3.86–3.91
Table 3.46 lists the heat of conversion for normal hydrogen to parahydrogen.
Figure 3.92 presents similar data graphically.
Table 3.47 gives the fixed points for hydrogen, and Tables 3.48 and 3.49 sum-
marize the fixed point properties of normal hydrogen and parahydrogen.
A catalyst is normally required to convert between the ortho and para forms of
hydrogen. Without a catalyst the ortho to para reaction proceeds only very slowly in
the liquid phase and not at all in the vapor phase. In the liquid phase, unconverted
normal liquid hydrogen will slowly convert to p-H2, releasing sufficient heat to
vaporize most of the liquid. Thus, ortho-to-para conversion must be incorporated
into the hydrogen liquefier so that the product liquid will consist almost entirely
of stable para-hydrogen.
To minimize the impact of heat removal due to ortho-to-para conversion on
the thermodynamic efficiency of the liquefier cycle, it is important to accomplish
ortho-to-para conversion at the highest possible temperature levels consistent with
thermodynamic equilibrium. Continuous conversion would be the most thermody-
namically efficient. In continuous conversion, ortho-to-para conversion occurs
simultaneously with cooling so that the hydrogen cooling path closely follows the
equilibrium curve. (Some offset from full equilibrium is necessary to provide some
driving force for conversion). Continuous conversion, however, requires a complex
integration of catalyst and heat exchange surface that is difficult to design into prac-
tical systems. A common design approach is to approximate continuous conversion
by using several simple adiabatic catalyst converters interdispersed with heat
exchange. Starting at around liquid nitrogen temperatures, hydrogen is partially
cooled, removed from the heat exchanger, and sent through an adiabatic converter
where the hydrogen partially rewarms due to the heat of conversion. Conversion to
parahydrogen is accomplished to the extent permitted by approaching near equili-
brium at the converter outlet temperature. The partially converted hydrogen is
returned to the heat exchanger, and the stepwise heat exchange-conversion process
is repeated until the product hydrogen from the final converter is at least 95% para-
hydrogen.
Two types of ortho–para catalysts are commercially available: an iron oxide
catalyst and a nickel silicate catalyst. The nickel silicate catalyst is the more active
of the two. Anywhere from 50 to a few hundred pounds of catalyst would be used
for each ton per day of liquefier plant capacity.
Most of the physical properties of hydrogen and deuterium, such as vapor
pressure, density of the liquid, and triple point temperature and pressure, are mildly
dependent upon the ortho–para composition. The transition from one form to the
other involves a substantial energy change, as shown earlier in Table 3.45.
Unfortunately, the uncatalyzed conversion of orthohydrogen to parahydrogen
is relatively slow, which leads to an important consideration in the liquefaction and

© 2005 by Marcel Dekker


5367-4 Flynn Ch03 R2 090804

Cryogenic Fluids 191

Figure 3.86 Interim T–S chart for parahydrogen.

storage of hydrogen that does not exist for other gases. As mentioned, equilibrium
hydrogen at room temperature consists of 75% orthohydrogen and 25% parahydro-
gen. At 20.39 K, however, the approximate temperature at which hydrogen boils at
atmospheric pressure, the equilibrium concentration, is 99.8% parahydrogen. If one
starts with equilibrium or normal hydrogen at room temperature and liquefies it, the
liquid orthohydrogen will slowly convert to parahydrogen. Since parahydrogen is

© 2005 by Marcel Dekker


5367-4 Flynn Ch03 R2 090804

192 Chapter 3

Figure 3.87 Temperature–entropy chart for parahydrogen.

the lower energy form, heat is liberated in the conversion process, and this is suffi-
cient to cause evaporation of nearly 70% of the hydrogen originally liquefied. Lique-
faction cycles, therefore, usually incorporate a catalyst through which the newly
liquefied hydrogen must pass. This causes the conversion to take place in the lique-
fier, where the heat can be absorbed conveniently, and eliminates the otherwise
gradual evaporation of the stored liquid from this cause.

© 2005 by Marcel Dekker


5367-4 Flynn Ch03 R2 090804

Cryogenic Fluids 193

Figure 3.88 Interim T–S chart for parahydrogen, upper temperatures.

If an ortho–para catalyst is not employed in the liquefier, it is possible to pro-


duce liquid normal hydrogen. The conversion from normal to equilibrium hydrogen
will then occur in the storage container. The heat of conversion is then unavoidably
present to cause evaporation of the stored product, even in a perfectly insulated
vessel. The simple derivation below gives the rate of this conversion and the effect
it has on the stored contents.

© 2005 by Marcel Dekker


5367-4 Flynn Ch03 R2 090804

194 Chapter 3

Figure 3.89 Parahydrogen temperature–entropy diagram, solid region.

The fractional rate of conversion is given by


dx
¼ kx2 ð3:25Þ
dt
where x is the ortho fraction at time t.
The mass rate of conversion is given by
dx
m ¼ kmx2 ð3:26Þ
dt
where m is the mass of liquid remaining in the container at time t.

© 2005 by Marcel Dekker


5367-4 Flynn Ch03 R2 090804

Cryogenic Fluids 195

Figure 3.90 Temperature–entropy chart for deuterium.

The rate of generation of heat is

dq dx
¼ Hm ¼ Hkmx2 ð3:27Þ
dt dt

where H is heat generated when converting unit mass from ortho- to parahydrogen.

© 2005 by Marcel Dekker


5367-4 Flynn Ch03 R2 090804

196 Chapter 3

Figure 3.91 Temperature–entropy chart for deuterium.

The rate of evaporation of liquid is

dm 1 dq Hk
¼ ¼ mx2 ð3:28Þ
dt L dt L
where L is the latent heat of vaporization per unit mass.

© 2005 by Marcel Dekker


5367-4 Flynn Ch03 R2 090804

Cryogenic Fluids 197

Table 3.46 Heat of Conversion from Normal Hydrogen to


Parahydrogen

Heat of conversion

Temp.(K) J=g cal=mol

10 527.139 253.9865
20 527.140 253.987
20.39 527.138 253.986
30 527.138 253.986
33.1 527.138 253.986
40 527.117 253.976
50 526.845 253.845
60 525.531 253.212
70 521.770 251.400
80 513.932 247.623
90 500.757 241.275
100 481.671 232.079
120 427.248 205.857
150 322.495 155.385
200 163.774 78.91
250 70.524 33.98
298.16 28.558 13.76
300 27.562 13.28

Figure 3.92 Heat of conversion for hydrogen from normal to para.

© 2005 by Marcel Dekker


5367-4 Flynn Ch03 R2 090804

198 Chapter 3

Table 3.47 Fixed Points for Hydrogen

p-H2 or e-H2 n-H2

Critical point
T ¼ 32.976  0.05 K (59.357 R) T ¼ 33.19 K (59.742 R)
P ¼ 12.759 atm (1.2928 MPa) P ¼ 12.98 atm (1.315 MPa)
r ¼ 15.59 mol=L (31.43 kg=m3) r ¼ 14.94 mol=L (30.12 kg=m3)
Normal boiling point
T ¼ 20.268 K (36.482 R) T ¼ 20.39 K (36.702 R)
>P ¼ 1 atm (0.101325 MPa) P ¼ 1 atm (0.101325 MPa)
rliq ¼ 35.11 mol=L (70.78 kg=m3) rliq ¼ 35.2 mol=L (71.0 kg=m3)
rvap ¼ 0.6636 mol=L (1.338 kg=m3) rvap ¼ 0.6604 mol=L (1.331 kg=m3)
Triple point
T ¼ 13.803 K (24.845 R) T ¼ 13.957 K (25.123 R)
P ¼ 0.0695 atm (0.00704 MPa) P ¼ 0.0711 atm (0.00720 MPa)
rsolid ¼ 42.91 mol=L (86.50 kg=m3) rsolid ¼ 43.01 mol=L (86.71 kg=m3)
rliq ¼ 38.21 mol=L (77.03 kg=m3) rliq ¼ 38.3 mol=L (77.2 kg=m3)
rvap ¼ 0.0623 mol=L (0.126 kg=m3) rvap ¼ 0.0644 mol=L (0.130 kg=m3)

Source: McCarty and Roder (1981).

Equation (3.28) gives the rate of variation of m with t. To solve this equation, it
is convenient to first express x as an explicit function of t, i.e., solve Eq. (3.25). The
general solution is 1=x ¼ kt þ A, where A is the constant of integration.
Starting with normal liquid in the dewar, x ¼ 0.75 when t ¼ 0, from which
A ¼ 1.33. then
1
¼ kt þ 1:33
x
or
1
x¼ ð3:29Þ
kt þ 1:33
Substitution of this value for x into Eq. (3.28) gives
!
dm kH m
¼
dt L ðkt þ 1:33Þ2
!
dm kH dt
¼
m L ðkt þ 1:33Þ2
 
kH 1
ln m ¼ þ In B ð3:30Þ
L kð1:33 þ ktÞ
where ln B is a constant of integration. Now write
m H
ln ¼ ð3:31Þ
B Lð1:33 þ ktÞ
or
 
m H
¼ exp ð3:31aÞ
B Lð1:33 þ ktÞ

© 2005 by Marcel Dekker


Table 3.48 Fixed Point Properties of Normal Hydrogen

Cryogenic Fluids
Triple point Normal boiling point Standard conditions

Property Solid Liquid Vapor Liquid Vapor Critical pointa STP(0 C) NTP(20 C)

Temperature (K) 13.957b 13.957b 13.957b 20.390b 20.390b 33.19b 273.15 293.15
Pressure (mmHg) 54.04 54.04 54.04 760 760 9865 760 760
Density (mol=cm3)  103 43.01 38.30 0.0644 35.20 0.6604 14.94 0.04460 0.04155
Specific volume (cm3=mol)  103 0.02325 0.026108 15.519 0.028409 1.5143 0.066949 22.423 24.066
Compressibility factor, Z ¼ PV=RT – 0.001621 0.9635 0.01698 0.9051 0.3191 1.00042 1.00049
Heats of fusion and vaporizationc (J=mol) 117.1 911.3 – 899.1 – 0 – –
Specific heat (J=(mol K)
Cs, at saturation 5.73 13.85 –46.94 18.91 –33.28 (Very large) – –
CP, at constant pressure – 13.23 21.22 19.70 24.60 (Very large) 28.59 28.89
CV, at constant volume – 9.53 12.52 11.60 13.2 (19.7) 20.30 20.40
Specific heat ratiod, g ¼ CP=CV – 1.388 1.695 1.698 1.863 (Large) 1.408 1.416
Enthalpye (J=mol) 321.6 438.7 1350.0 548.3 1447.4 1164 7749.2 8324.1
Internal energye (J=mol) 317.9d 435.0 1234.8 545.7 1294.0 – 5477.1 5885.4
Entropye (J=mol k) 20.3 28.7 93.6 34.92 78.94 54.57 139.59 141.62
Velocity of sound (m=s) – 1282 307 1101 357 – 1246 1294
Viscosity, m – 0.026 0.00074 0.0132 0.0011 (0.0035) 0.00839 0.00881
N s=m2  103
centipoise – 0.026 0.00074 0.0132 0.0011 (0.0035) 0.00839 0.00881
Thermal conductivity, k(mW(cm K)) 9.0 0.73 0.124 0.99 0.169 –f 1.740 1.838
Prandtl number, Npr ¼ mCp=k – 2.24 0.623 1.29 0.809 – 0.6849 0.6848
Dielectric constant, e 1.287 1.253 1.00039 1.231 1.0040 1.0937 1.000271 1.000253
pffiffi
Index of refractiong,h, n ¼ e 1.134 1.119 1.000196 1.1093 1.0020 1.0458 1.000136 1.000126
Surface tension (N=m)  103 – 3.00 – 1.94 – 0 – –
Equiv. vol.=vol. liquid at NBP 0.8184 0.9190 546.3 1 53.30 2.357 789.3 847.1

Gas constant: R ¼ 8.31434 J(mol K); molecular weight ¼ 2.01594; mole ¼ gram mole.
a
Values in parentheses are estimates.
b
These temperatures are based on the IPTS-1968 temperature scale. To compare with the corresponding temperatures based on the NBS-1955 temperature scale, 0.01 K should
be subtracted; i.e., use 13.947, 20.380, and 33.18, respectively, for the triple point, normal boiling point, and critical point when determining property differences between

199
normal and parahydrogen.
c
Heats of fusion and vaporization calculated from enthalpy differences.
d
Calculated from property values given in this table.
e
Base point (zero values) for enthalpy, internal energy, and entropy are 0 K for the ideal gas at 0.101325 Mpa (1 atm) pressure.
f
Anomalously large.
g
Long wavelengths.
h
Index of refraction calculated from dielectric constant data.

© 2005 by Marcel Dekker


Table 3.49 Fixed Point Properties of Parahydrogen

200
Triple point Normal boiling point Standard conditions

Property Solid Liquid Vapor Liquid Vapor Critical pointa STP(0 C) NTP(20 C)

Temperature (K) 13.803 13.803 13.803 20.268 20.268 32.976 273.15 293.15
Pressure (mmHg) 52.82 52.82 52.85 760 760 9696.8 760 760
Density (mol=cm3)  103 42.91 38.207 0.0623 35.11 0.6636 15.59 0.5459 0.04155
Specific volume (cm3=mol)  103 0.02330 0.026173 16.057 0.028482 1.5069 0.064144 22.425 24.069
Compressibility factor, Z ¼ PV=RT – 0.001606 0.9850 0.01712 0.9061 0.3025 1.0005 1.0006
Heats of fusion and vaporizationb (J=mol) 117.5 905.5 – 898.3 – 0 – –
Specific heat (J=(mol K)
Cs at saturation 5.73 13.85 46.94 18.91 33.28 (Very large) – –
Cp, at constant pressure – 13.13 21.20 19.53 24.50 (Very large) 30.35 30.02
CV, at constant volume – 9.50 12.52 11.57 13.11 19.7 21.87 21.70
Specific heat ratioc g ¼ CP=CV – 1.382 1.693 1.688 1.869 (Large) 1.388 1.383
Enthalpyd (J=mol) 740.2 622.7 282.8 516.6 381.7 77.6 7656.6 8260.6
Internal energyd (J=mol) 740.4c 622.9 169.8 519.5 229.0 5.7 5384.5 5822.0
Entropyd (J=(mol K) 1.49 10.00 75.63 16.08 60.41 35.4 127.77 129.90
Velocity of sound (m=s) – 1273 305 1093 355 350 1246 1294
Viscosity,m – 0.026 0.00074 0.0132 0.0011 0.0035 0.00839 0.00881
N s=m2  103
centipoise – 0.026 0.00074 0.0132 0.0011 0.0035 0.00839 0.00881
Thermal conductivity,k(mW(cm K)) 9.0 0.73 0.124 0.99 0.169 –e 1.841 1.914
Prandtl number, Npr ¼ mCp=k – 2.24 0.623 1.29 0.809 0.6866 0.6855
Dielectric constant, e 1.286 1.252 1.00038 1.230 1.0040 1.098 1.00027 1.00026
pffiffi
Index of refraction f,g, n ¼ e 1.134 1.119 1.00019 1.109 1.0020 1.048 1.00013 1.00012
Surface tension (N=m)  103 – 2.99 – 1.93 – 0 – –
Equiv. vol.=vol. liquid at NBP 0.8181 0.9190 563.8 1 52.91 2.252 787.4 845.1

Chapter 3
Gas constant: R ¼ 8.31434 J(mol K); molecular weight ¼ 2.01594; mole ¼ gram mole.
a
Values in parentheses are estimates.
b
Heats of vaporization calculated from enthalpy differences.
c
Calculated from property values given in this table.
d
Base point (zero values) for enthalpy, internal energy, and entropy are 0 K for the ideal gas at 0.101325 MPa (1 atm) pressure.
e
Anomalously large.
f
Long wavelengths.
g
Index of refraction calculated from dielectric constant data.

© 2005 by Marcel Dekker


5367-4 Flynn Ch03 R2 090804

Cryogenic Fluids 201

when t ¼ 0, m ¼ mo, so
 
mo 3H mo
¼ exp or B¼ ð3:32Þ
B 4L expð3H=4LÞ
The solution then becomes
   
m 3H H
exp ¼ exp ð3:33Þ
mo 4L Lð1:33 þ ktÞ
or
 
m H 3H
¼ exp  ð3:33aÞ
mo Lð1:33 þ ktÞ 4L
L ¼ 906.09 J=mol (216.56 cal=mol) for n-H2 at 20 K (36 R) and 894.87 J=mol
(213.88 cal=mol) for 99.8% p-H2 at 20 K. 890 J=mol (215 cal=mol) is a reasonable
average value. The value for H is 1416.19 J=mol (338.65 cal=mol), and the value
for the rate constant k is 0.0114=hr1.
These numerical substitutions reduce Eq. (3.33) to the form
 
m 1:57
¼ exp  1:18 ð3:34Þ
mo ð1:33 þ 0:0114tÞ
or
m 1:57
ln ¼  1:18 ð3:34aÞ
mo 1:33 þ 0:0114t
If the original composition of the liquid is not 0.75 orthohydrogen at t ¼ 0, then
a new constant of integration based on the starting composition can be evaluated for
Eq. (3.29). The new expression for m=mo then follows directly as outlined above, and
Fig. 3.93 shows the results.

3.7.3. Properties of Hydrogen


Physical constants for five isotopic forms of hydrogen are presented in Table 3.50.
a. Melting Curve. The melting curve is the boundary between the solid and
liquid regions in a phase diagram. Range of values (kg=m3) for parahydrogen:

Triple point 10.1325 Mpa (100 atm) 30.3975 MPa (300 atm)

Liquid Solid Fluid Solid Fluid Solid

77.0 86.9 81.8 89.9 88.7 96.0

Uncertainty is estimated to be 0.1% for the liquid phase and 0.5% for the solid phase.
b. Heat of Vaporization. Range of values (J=g):

Triple point Boiling point

Para 448.2 445.5


Normal 449.1 445.6

© 2005 by Marcel Dekker


5367-4 Flynn Ch03 R2 090804

202 Chapter 3

Figure 3.93 Fraction of liquid hydrogen evaporated due to ortho–para conversion as a func-
tion of time.

Uncertainty is estimated to be 1.24 J=g for p-H2 and 2.5 J=g for n-H2.
The heat of vaporization for parahydrogen and normal hydrogen is tabulated
in Table 3.51 and graphed in Figs. 3.94 and 3.95.
c. Vapor Pressure. The vapor pressure is the pressure, as a function of
temperature, of a liquid in equilibrium with its own vapor.

Table 3.50 Physical Constants of the Isotopic Forms of Hydrogen

20.39 20.39
Normal Equilibrium Normal Equilibrium
hydrogen, hydrogen, deuterium, deuterium, Hydrogen
75% o-H2 0.21% o-H2 66.67% o-D2 97.8% o-D2 deuteride

Triple point
temp., K 13.96 13.81 18.72 18.69 16.60
Triple point pressure
atm 0.07105 0.0695 0.1691 0.1691 0.122
kPa 0.07197 0.0704 0.1713 0.1713 0.124
Normal boiling
point, K 20.39 20.27 23.57 23.53 22.13
Critical temp., K 33.19 33.1 38.3 35.9
Critical pressure
atm 12.98 12.8 16.3 14.6
kPa 13.15 13.1 16.5 14.8
Critical volume,
cm3=mol 66.95 65.5 60.3 62.8

© 2005 by Marcel Dekker


5367-4 Flynn Ch03 R2 090804

Cryogenic Fluids 203

Table 3.51 Heat of Vaporization (hv) of Normal


Hydrogen

Pressure (atm) Temperature (K) hv (cal=g)

0.12 15.0 108.6


0.6 17.8 106.7
0.8 19.7 106.0
1.0 20.5 105.5
1.5 22.0 103.3
2.0 23.0 102.0
3.0 24.75 98.0
4.0 26.2 94.3
5.0 27.3 89.0
6.0 28.3 84.0
8.0 30.0 71.2
10.0 31.3 57.1
12.0 32.6 33.0
12.98 33.19 0

Source: Wooley et al. (1948); Simon and Lange (1923).

Figure 3.94 Heat of vaporization of parahydrogen.

© 2005 by Marcel Dekker


5367-4 Flynn Ch03 R2 090804

204 Chapter 3

Figure 3.95 Latent heat of vaporization of normal hydrogen.

Parahydrogen:
ðnpÞ
ln ¼ n1 x þ n2 x2 þ n3 x3 þ n4 ð1  xÞn5 ð3:35Þ
ðnpt Þ

where vpt ¼ 0.007042 Mpa, n1 ¼ 3.05300134164, n2 ¼ 2.80810925813, n3 ¼


0.655461216567, n4 ¼ 1.59514439374, n5 ¼ 1.5814454428, and

1  Tt =T

1  Tt =Tc

with

Tt ¼ 13:8 K; TC ¼ 32:938 K
Figures 3.96 and 3.97 present curves for the vapor pressure of p-H2 below and
above p ¼ 1 atm, respectively.
Normal hydrogen:

log Pðkg=cm2 Þ ¼ A þ B=T þ C log T ð3:36Þ

© 2005 by Marcel Dekker


5367-4 Flynn Ch03 R2 090804

Cryogenic Fluids 205

Figure 3.96 Vapor pressure of parahydrogen below 0.101325 MPa (1 atm).

where
A ¼ 1:55447; B ¼ 31:875; C ¼ 2:39188

and T is in kelvins.
Table 3.52 presents vapor pressure values for the isotopic forms of hydrogen.
The vapor pressure equations of Table 3.52 agree with the experimental data
for hydrogen to within the experimental error over most of the liquid range to the
critical point. The values for deuterium and hydrogen deuteride should not be used
at pressures much above 1 atm.
Figures 3.98 and 3.99 present vapor pressure curves for n-H2.
d. Virial Coefficients. The virial coefficients are commonly defined in two
ways as follows:

P ¼ RT r½1 þ BðTÞr þ CðTÞr2 þ   

and

PV ¼ RT þ B0 ðTÞP þ C 0 ðTÞP2

The first two terms of these series may be inverted by the following:

CðTÞ  BðTÞ2
BðTÞ ¼ B0 ðTÞ and C 0 ðTÞ ¼ ð3:37Þ
RT

© 2005 by Marcel Dekker


5367-4 Flynn Ch03 R2 090804

206 Chapter 3

Figure 3.97 Vapor pressure of parahydrogen above 0.101325 MPa (1 atm).

where P is the pressure, T is the temperature, r is the density, V is the 1=r, and R is
the gas constant. B(T ), B0 (T ) andC(T ), C0 (T ) are virial coefficients of a power series
expansion in density and pressure, respectively.
Both series are theoretically infinite in length; however, the coefficients beyond
the first two [B(T ) and C(T ) ] are of less interest because of their complexity. The
temperature at which B(T) ¼ 0 is called the Boyle point. Either of the above two

Table 3.52 Vapor Pressure of the Isotopic Forms of Hydrogen

Material State A B C

Normal hydrogen, 75% o-H2 Liquid 4.66687 44.9569 0.020537


Solid 4.56488 47.2059 0.03939
20.39 K Equilibrium hydrogen, 0.21% o-D2 Liquid 4.64392 44.3450 0.02093
Solid 4.62438 47.0172 0.03635
Normal deuterium, 66.67% o-D2 Liquid 4.7312 58.4619 0.02671
Solid 5.1626 68.0782 0.03110
20.39 K Equilibrium deuterium, 97.8% o-D2 Liquid 4.7367 58.4440 0.02670
Solid 5.1625 67.9119 0.03102
Hydrogen deuteride Solid 5.04964 55.2495 0.01479
Liquid 4.70260 56.7154 0.04101

© 2005 by Marcel Dekker


5367-4 Flynn Ch03 R2 090804

Cryogenic Fluids 207

equations is adequate to describe the PVT surface for densities up to about one-half
the critical density.
Range of values:

Unit 20 K 100 K 200 K 300 K


3
B(T) cm =mol 146.7 2.51 10.73 14.48
C(T) (cm3=mol)2 1503 608.6 421.7 343.8

Boyle point (B ¼ 0):112.4 K.

The uncertainty of B is probably a maximum of 5% at the highest and lowest tem-


peratures. The uncertainty of C is a minimum of 5% between 55 and 100 K and as
much as 20% below critical temperature.
Values for B and C for parahydrogen are listed in Table 3.53.
Table 3.54 gives values for converting between specific volume and density.
e. Joule–Thomson Coefficient. The Joule–Thomson coefficient m is defined as
   
@H @T
m ¼ Cp  1 ¼ ð3:38Þ
@P T @P H
The sign of m indicates whether the expansion of a gas will cause an increase or
decrease in its temperature. If m is positive, the expanding gas will be cooled. The
locus of points where m ¼ 0 is called the Joule–Thomson (J–T) inversion curve. Data

Figure 3.98 Vapor pressure of normal hydrogen below 0.101325 MPa (1 atm).

© 2005 by Marcel Dekker


5367-4 Flynn Ch03 R2 090804

208 Chapter 3

Figure 3.99 Vapor pressure of normal hydrogen above 0.101325 MPa (1 atm).

for the J–T curve for parahydrogen are given in Table 3.55. The uncertainty in the
data is estimated to be 5%.
Figure 3.100 shows the J–T inversion curve for parahydrogen.
f. Dielectric Constant. The dielectric constant for fluid parahydrogen may be
calculated from
e1
¼ Ar þ Br2 þ Cr3 ð3:39Þ
eþ2
where A ¼ 0.99575, B ¼ 0.09069, C ¼ 1.1227, and e is the dielectric constant. The den-
sity, r, must be in units of grams per cubic centimeter. The equation is valid over the
range of the tables and may be extrapolated in the fluid phase with reasonable results.
Range of values (dimensionless) :

Triple point Boiling point

Liquid Vapor Solid Liquid Vapor Critical point 300 K, 1 atm

1.252 1.0004 1.285 1.229 1.004 1.0980 1.00025

© 2005 by Marcel Dekker


5367-4 Flynn Ch03 R2 090804

Cryogenic Fluids 209

Table 3.53 Second and Third Virial Coefficients for Parahydrogen

Temp. (K) B (cm3=mol) C (cm3=mol)2

14 2.372  102 6.713  104


16 1.992  102 2.265  104
18 1.698  102 7.121  103
20 1.467  102 1.503  103
22 1.281  102 5.656  102
24 1.129  102 1.323  103
26 1.003  102 1.580  103
28 8.969  101 1.640  103
30 8.066  101 1.615  103
32 7.290  101 1.550  103
34 6.615  101 1.467  103
36 6.025  101 1.376  103
38 5.504  101 1.290  103
40 5.042  101 1.214  103
45 4.086  101 1.066  103
50 3.343  101 9.638  102
60 2.264  101 8.351  102
70 1.522  101 7.510  102
80 9.824  101 6.878  102
90 5.725  101 6.405  102
100 2.510  101 6.085  102
120 2.144  101 5.491  102
140 5.342  101 5.063  102
160 7.654  101 4.726  102
180 9.388  101 4.449  102
200 1.073  101 4.217  102
250 1.300  101 3.768  102
300 1.438  101 3.438  102
400 1.589  101 2.976  102
500 1.660  101 2.661  102

The uncertainty of e – 1 is estimated to be no greater than 0.1% for the fluid


phase and 0.2% for the solid phase. Values of the dielectric constant are listed in
Table 3.56.
Symbols: rsat ¼ saturation pressure or vapor pressure, r ¼ density, CP ¼ specific
heat at constant pressure, m ¼ viscosity, k ¼ thermal conductivity, hfg ¼ heat of
vaporization, Npr ¼ mCP=K ¼ Prandtl number, sL ¼ surface tension, e ¼ dielectric
constant.

3.7.4. The Hydrogen Density Problem


Hydrogen has the lowest liquid density of any known substance. This is an in-
escapable fact of liquid hydrogen behaving as a quantum fluid. This is true because
liquid hydrogen possesses a high quantum mechanical zero-point energy relative to
its classical thermal energy. According to classical theory, at absolute zero tempera-
ture the particles of matter should be in static equilibrium with one another. Since
they then have no thermal energy, perfect static balance is supposed to exist between
the electromagnetic attractive and repulsive forces of the atoms.

© 2005 by Marcel Dekker


5367-4 Flynn Ch03 R2 090804

210 Chapter 3

Table 3.54 Conversion Table for Specific Volumes to


Density, for Hydrogen

Specific volume

cm3=mol ft3=lb Density (lb=ft3)

26 0.207 4.84
26.5 0.211 4.75
27 0.215 4.66
27.5 0.219 4.58
28 0.222 4.49
28.5 0.226 4.42
29 0.230 4.34
29.5 0.234 4.27
30 0.238 4.20
31 0.246 4.06
32 0.254 3.93
33 0.262 3.81
34 0.270 3.70
35 0.278 3.60
36 0.286 3.50
37 0.294 3.40
38 0.302 3.31
39 0.310 3.23
40 0.318 3.15
42 0.334 3.00
44 0.350 2.86
46 0.365 2.74
48 0.381 2.62
50 0.397 2.52
55 0.437 2.29
60 0.477 2.10
65 0.516 1.94
70 0.556 1.80

Above 0 K, however, all matter has thermal energy in the form of rapid
random motion of its atoms, and the balance of forces among the particles becomes
dynamic rather than static. In cooling matter slowly to 0 K, then, classical theory
would predict the loss of thermal energy of the material through the loss of the
kinetic energy of its atoms until at 0 K there would exist perfect motionless order.
Quantum theory, on the other hand, shows that each atom has an irreducible
minimum of kinetic energy amounting to (1=2) hv, where h is Planck’s constant,
6.6  1027 erg sec, and v is the frequency of oscillation of the atom. Even at 0 K,
when a substance has lost all its thermal energy, it will still have this zero-point
energy.
This amount is not large, and in most cases it is effectively inundated by the
thermal energy of matter at higher temperatures. At very low temperatures, however,
the zero-point energy becomes a significant fraction of the total energy of some
substances.
Solid hydrogen, for example, has a zero-point energy of about 200 cal=mol,
which counteracts about 50% of its computed lattice energy of 400 cal=mol.

© 2005 by Marcel Dekker


5367-4 Flynn Ch03 R2 090804

Cryogenic Fluids 211

Table 3.55 Joule–Thomson Inversion Curve Data for Parahydrogen

Temperature Pressure Density

K 
R MPa atm psia mol=cm3  103 lbm=ft3

28 50.4 1.000 9.87 145.1 30.06 3.783


29 52.2 1.525 15.05 221.2 29.90 3.763
30 54.0 2.035 20.08 295.1 29.73 3.742
31 55.8 2.534 25.01 367.6 29.56 3.720
32 57.6 3.025 29.85 438.7 29.40 3.700
34 61.2 3.968 39.16 575.5 29.05 3.656
36 64.8 4.870 48.06 706.3 28.70 3.612
40 72.0 6.545 64.59 949.2 27.99 3.523
50 90.0 10.02 98.93 1454 26.16 3.292
60 108.0 12.60 124.4 1828 24.30 3.058
80 144.0 15.55 153.5 2256 20.58 2.590
100 180.0 16.35 161.4 2372 17.04 2.145
120 216.0 16.42 162.1 2353 14.12 1.777
140 252.0 14.24 140.5 2064 10.86 1.367
160 288.0 10.36 102.2 1502 7.176 0.9031
180 324.0 5.165 50.97 749.1 3.321 0.4179
200 360.0 0.0547 0.54 8.6 0.036 0.0045

The measured heat of sublimation of hydrogen is therefore only 400200 ¼


200 cal=mol. The zero-point energy acts as though it were additional thermal energy
and effectively counteracts part of the attractive force between molecules of hydro-
gen. The practical result is that solid hydrogen melts very easily. Indeed, the thermal
properties of solid hydrogen resemble those of liquid helium more than they do those
of liquid hydrogen. Solid hydrogen melts readily, and liquid hydrogen vaporizes very
easily.
Figure 3.101 is the familiar plot of intermolecular potential energy vs. intera-
tomic separation distance. Above the horizontal axis, the coulombic forces domi-
nate, causing a repulsion among the molecules. Below the abscissa, the
gravitational attractive forces between the molecules dominate. The lower dashed
curve is the classical energy curve representing the balance between repulsive and
attractive forces among molecules vs. their separation. The upper dashed curve
shows the zero-point energy predicted by quantum mechanics. The solid curve is
the sum of the classical energy and quantum energy forces vs. molecular separation.
The average interatomic spacing is the sum of the classical and quantum forces and
is considerably larger than if the zero-point energy were negligible.
The effect of the zero-point energy is to shift the curve to the right in Fig. 3.101
R0 0 is the molecular separation for a classical fluid. R0 is the molecular separation
that results from adding the zero-point energy to the classical energy. Shifting R0
to the right inflates the distance between the molecules and therefore decreases the
density of the liquid. This is an inescapable fact of life for liquid hydrogen and
accounts for the extremely low liquid density of 70 g=L or a specific gravity of
only 7%.
The atoms in liquid hydrogen do not pack well. Indeed, there is more hydrogen
in a gallon of water than there is in a gallon of pure liquid hydrogen. Methane is a
much better carrier of hydrogen than is hydrogen, as shown in Table 3.57.

© 2005 by Marcel Dekker


5367-4 Flynn Ch03 R2 090804

212 Chapter 3

Figure 3.100 Joule–Thomson inversion curve for parahydrogen.

Table 3.56 Dielectric Constant for Parahydrogen

Density (g=cm3) Dielectric constant, e

0.005 1.01515
0.010 1.03046
0.015 1.04594
0.020 1.06158
0.025 1.07739
0.030 1.09336
0.035 1.10950
0.040 1.12580
0.045 1.14226
0.050 1.15889
0.055 1.17569
0.060 1.19265
0.065 1.20977
0.070 1.22705
0.075 1.24449
0.080 1.26210

© 2005 by Marcel Dekker


5367-4 Flynn Ch03 R2 090804

Cryogenic Fluids 213

Figure 3.101 The effect of zero-point energy on atomic separation.

Another effect of the zero—point energy is to reduce the depth of the potential
well in Fig. 3.101. The ‘‘well’’ region below the horizontal axis actually represents the
liquid phase of a fluid. It is only here that the fluid can exist in two different densities
(interatomic separations) at the same potential energy (temperature). The potential
well therefore represents the liquid–vapor equilibrium region of a fluid. Reducing
the depth of the potential well reduces the temperature range over which vapor–
liquid equilibrium is possible. Therefore, the depth of the potential well is propor-
tional to the heat of vaporization (or sublimation) of the fluid. Hence, the low heat
of vaporization of liquid hydrogen is a macroscopic manifestation of the zero-point
energy of liquid hydrogen.
We shall see shortly that the zero-point energy is manifested even more promi-
nently in liquid helium. It should be noted that neon also exhibits strong quantum
mechanical effects. That is why these three fluids—hydrogen, helium, and neon—
are referred to as the quantum fluids. This is also the reason the PVT properties
of these three fluids are omitted from the usual generalized correlations of state.
These equations depend upon the similarity among fluids, especially the rule that
fluids equally distant from their critical points behave similarly. That is to say, at
the same reduced temperature, TR ¼ T=Tc, and the same reduced pressure,
PR ¼ P=PC, the fluids will have the same compressibility factor, Z ¼ PV=RT.
Helium, hydrogen, and neon do not follow the rules of normal fluids and do not
fit well the generalized equations of state. They are quantum fluids instead.

Table 3.57 How Much Hydrogen is There in a Gallon of Liquid Hydrogen?

Fluid Density at NBP (g=cm3) mol=cm3 at NBP moles H2=cm3 ratio to H2

H2 0.0707 0.03507 0.03507 1.00


CH4 0.5110 0.03194 0.063875 1.82
H2O 1.0000 0.05560 0.05560 1.58

© 2005 by Marcel Dekker


5367-4 Flynn Ch03 R2 090804

214 Chapter 3

Other Properties. Figures 3.102–3.113 give the dielectric constant, specific


heats, density, melting curve, surface tension, vapor pressure, differences in specific
heat between normal and para-differences between the thermal conductivity of
normal and para- of hydrogen taken from the WADD report mentioned before
(Johnson, 1960).

3.7.5. Uses of Hydrogen


More than 97% of current domestic hydrogen use is in chemical manufacturing and
petroleum refining. The other hydrogen uses are often important in their specific
industries.
a. Petroleum Refining. Hydrogen is used to treat and upgrade oils in hydro-
treating and hydrocracking processes in petroleum refining. Hydrotreating, the
reacting of hydrogen with refinery fractions to desulfurize feedstocks, to hydrogenate
olefins, and to treat lube oils and kerosene-type jet fuels, has grown rapidly with the
introduction of catalytic reforming and the resulting supply of low-cost hydrogen.
Catalytic reforming produces around 125 m3 of hydrogen for every cubic meter of
oil processed (700 ft3=bl). The amount of hydrogen required for hydrotreating ranges
from 9 m3 per cubic meter of straight-run naphthas (50 ft3=bl) to more than 15 times
as much for some coker distillates. The average is estimated to be around 44.5 m3 per
cubic meter of feed (250 ft3=bl).
b. Chemicals. Between 1982 and 2266 m3 (70,000–80,000 ft3) of hydrogen is
required to produce 907 kg (1 ton) of synthetic anhydrous ammonia, including the
amount that does not react to form ammonia.
The making of synthetic methanol requires about 2.2 m3 of hydrogen per kilo-
gram (36 ft3=lb).

Figure 3.102 Dielectric constant of saturated parahydrogen vapor.

© 2005 by Marcel Dekker


5367-4 Flynn Ch03 R2 090804

Cryogenic Fluids 215

Figure 3.103 Specific heat (CP) of parahydrogen at constant pressure.

Cyclohexane is directly recoverable from petroleum and is also made by the


catalytic hydrogenation of benzene. The benzene route for cyclohexane production,
representing 96% of cyclohexane capacity in 1972, uses a minimum of 814 m3
(28,750 ft3) of hydrogen per 907 kg (ton) of cyclohexane produced.

Figure 3.104 Specific heat CP of normal hydrogen at constant pressure.

© 2005 by Marcel Dekker


5367-4 Flynn Ch03 R2 090804

216 Chapter 3

Figure 3.105 Specific heat (CV) of parahydrogen at constant volume.

Aniline is produced from benzene by nitration and then catalytic hydrogena-


tion requiring a minimum of 0.81 m3 of hydrogen per kilogram of product (13 ft3=lb).
Naphthalene production was coal-based until the early 1960s when petroleum-
based processes were developed. Today, around 40% of naphthalene production is

Figure 3.106 Specific heat CV of normal hydrogen at constant volume.

© 2005 by Marcel Dekker


5367-4 Flynn Ch03 R2 090804

Cryogenic Fluids 217

Figure 3.107 Melting line for parahydrogen, critical point and above.

from petroleum. Hydrogen is used for desulfurization and dealkylation in a hydro-


gen atmosphere. Total hydrogen use in naphthalene production is estimated to be
around 1.87 m3 per kilogram of naphthalene (30 ft3=lb).
Other hydrogen uses in the chemical industry include the production of a vari-
ety of chemicals such as hydrogen chloride, aldehyde, caprolactam, ethylene, hydro-
gen peroxide, and styrene.
c. Hydrogenation of Fats and Oils. Hydrogen plays a major role in the hard-
ening of vegetable and fish oils used in the production of margarine, lard, shortening,
and cooking oils.

© 2005 by Marcel Dekker


5367-4 Flynn Ch03 R2 090804

218 Chapter 3

Figure 3.108 Melting line from triple point to critical point pressure for parahydrogen.

d. Aerospace. A significant amount of the liquid hydrogen produced, perhaps


half, is used by the aerospace industry for rocket propellant and rocket engine com-
ponent testing. Hydrogen is also used in spacecraft fuel cells.
e. Metallurgy. Hydrogen is used for direct reduction of ores, heat treatments,
welding, and in the production of pure metals. The major present use is in the direct
reduction of iron oxide ores to sponge iron, which uses around 566 m3 of hydrogen
per 907 kg iron produced (20,000 ft3=ton).
f. Electricity and Electronics. Some public electric utilities require hydrogen
gas as a coolant for large generators, motors, and frequency changers. In electronics,

© 2005 by Marcel Dekker


5367-4 Flynn Ch03 R2 090804

Cryogenic Fluids 219

Figure 3.109 Surface tension of liquid parahydrogen.

hydrogen is used as a reducing atmosphere and for brazing of metals and silicon
compounds for solid-state components, vacuum tubes, light bulbs, and crystal grow-
ing.
g. Other Uses. Hydrogen use in other areas is difficult to quantify. These use
areas include glass manufacture via the float glass process, oxygen-hydrogen glass
cutting, artificial gem production, refrigeration, pharmaceuticals, nuclear fuel pro-
cessing, and nuclear research using liquid hydrogen bubble chambers.

3.8. Helium
Evidence of the existence of helium (He) was first obtained by Janssen during the
solar eclipse of 1868 when he detected a new line in the solar spectrum; Lockyer
and Frankland suggested the name helium for the new element. In 1895, Ramsay
discovered helium in the uranium mineral clevite, and it was independently
discovered in clevite by the Swedish chemists Cleve and Langlet about the same
time.
Onnes first liquefied helium in 1908 in his laboratory at Leiden, Netherlands,
by using liquid hydrogen precooling in a Joule–Thomson liquefier. Liquid helium
was not available outside of Leiden until 1923.
In 1920, Aston discovered that, in addition to the common isotope helium-4,
there existed a very rare isotope helium-3. The helium-4 atom consists of two elec-
trons orbiting a nucleus of two protons plus two neutrons. The helium-4 nucleus
is the alpha particle associated with radioactive decay and other atomic processes.
The helium-3 atom consists of two electrons orbiting a nucleus of two protons plus
one neutron.

© 2005 by Marcel Dekker


5367-4 Flynn Ch03 R2 090804

220 Chapter 3

Figure 3.110 Surface tension of liquid normal hydrogen.

Helium-4 is by far the more common of the two isotopes. Ordinary helium gas
contains about 1.3  104% helium-3, so that when we speak of helium or liquid
helium, we normally are referring to helium-4 (molecular weight 4.0026). Liquid
helium-4 has a normal boiling point of 4.216 K (7.57 R) and a density at the normal
boiling point of 124.96 kg=m3 (7.8 lbm=ft3), or about one-eighth that of water.
The physical constants for helium are shown in Table 3.58.

3.8.1. Helium Phase Diagram


Helium, like hydrogen, possesses a high quantum mechanical zero-point energy rela-
tive to its classical thermal energy. As we mentioned earlier, solid hydrogen has a
zero-point energy of about 200 cal, which counteracts about 50% of its computed lat-
tice energy of 400 cal=mol. The zero-point energy contribution in helium is even
more dramatic.
The zero-point energy of solid helium near 0 K, about 50 cal=mol, counteracts
about 80% of the calculated lattice energy of 62 cal=mol, and the zero point energy of
liquid helium is about 80% of the cohesive energy.
The average interatomic spacing of liquid helium is determined by a balance of
the zero-point energy and the classical energy forces. Hence, the average interatomic
spacing of helium is also greatly affected by the zero-point energy. The density of
liquid helium is more gas-like than liquid-like due to the high zero-point energy.
One particularly gas-like property is the high compressibility of liquid helium. At
a pressure of 20,000 atm, helium-4 occupies only two-fifths of its volume at 1 atm.
The zero-point energy also reduces the binding energy between atoms in the
liquid state and hence reduces the heat of vaporization of the liquid. In most liquids,

© 2005 by Marcel Dekker


5367-4 Flynn Ch03 R2 090804

Cryogenic Fluids 221

Figure 3.111 Vapor pressure of normal and para-hydrogen below the triple point.

the entropy difference between the liquid and gaseous states (equal to the heat of
vaporization divided by the temperature) is about 80 cal=(mol K) at the normal boil-
ing point (Trouton’s rule). In liquid helium, the influence of zero-point energy causes
this number to drop to about 20 cal=mol. This zero-point energy accounts for the
very low heat of vaporization of liquid helium. It is as if the zero-point energy
had already started the pot to boil.
The large zero-point energy of liquid helium has a profound and totally unex-
pected effect on its phase diagram.
Figure 3.114 shows the phase diagram of a normal fluid on the left and the
phase diagram of helium on the right. The normal diagram shows that there are
three phases, solid, liquid, and vapor; that these are separated by the melting, vapor
pressure, and sublimation curves; that there is a critical point where the vapor and
liquid phases are no longer distinguishable; and that there is a solid–liquid–vapor
coexistence point, the triple point.

© 2005 by Marcel Dekker


5367-4 Flynn Ch03 R2 090804

222 Chapter 3

Figure 3.112 Difference in the specific heat of para- and normal hydrogen.

Figure 3.113 Difference in the thermal conductivity of para- and normal hydrogen.

© 2005 by Marcel Dekker


5367-4 Flynn Ch03 R2 090804

Cryogenic Fluids 223

Table 3.58 Physical Constants for Helium

Helium
Chemical Symbol: He
CAS Registry Number: 7440-59-7
DOT Classification: Nonflammable gas
DOT Label: Nonflammable gas
Transport Canada Classification: 2.2
UN Number: UN 1046 (compressed gas); UN 1963 (refrigerated liquid)

US Units SI Units

Physical constants
Chemical formula: He He
Molecular weight 4.00 4.00
Density of the gas at 70 F (21.1 C) and 1 atm 0.0103 lb=ft3 0.165 kg=m3
Specific gravity of the gas at 70 F (21.1 C) 0.138 0.138
and 1 atm
Specific volume of the gas at 70 F (21.1 C) 97.09 ft3=lb 6.061 m3=kg
and 1 atm
Density of the liquid at boiling point and 1 atm 7.802 lb=ft3 124.98 kg=m3
Boiling point at 1 atm –452.1 F –268.9 C
Melting point at 1 atm None None
Critical temperature –450.3 F –267.9 C
Critical pressure 33.0 psia 227 kPa abs
Critical density 4.347 lb=ft3 69.64 kg=m3
Triple point None None
Latent heat of vaporization at boiling point 8.72 Btu=lb 20.28 kJ=kg
and 1 atm
Latent heat of fusion at triple point No T.P. No T.P.
Specific heat of the gas at 70 F (21.1 C)
and 1 atm
CP 1.24 Btu=(lb)( F) 5.19 kJ=(kg)( C)
CV 0.745 Btu=(lb)( F) 3.121 kJ=(kg)( C)
Ratio of specific heats (CP=CV) 1.66 1.66
Solubility in water, vol=vol at 32 F (0 C) 0.0094 0.0094
Weight of the liquid at boiling point 1.043 lb=gal 125.0 kg=m3

The helium phase diagram (Fig. 3.114) exhibits many of the characteristics dis-
played by normal fluids: the existence of a critical point on the liquid–vapor curve,
for instance. One difference between helium and all other substances is the shape of
the solid–liquid (melting) curve. For other substances, this line intersects the liquid–
vapor line at some point, the triple point. For helium, there is no temperature
and pressure at which the solid, liquid, and vapor phases of helium can exist in
equilibrium. Helium has no triple point. Linear extrapolation of the solid–liquid line
would place a triple point at about 1 K. However, from about 2 K down, this extra-
polated line deviates from its higher temperature linear behavior and approaches 0 K
horizontally at a value of 25 atm. This means that helium can never be solidified at
atmospheric pressure merely by reducing its temperature, as is the case with all other
substances. In addition, it is necessary to subject the helium to a pressure of 25 atm.
This is a direct result of the quantum mechanical zero-point energy of helium.

© 2005 by Marcel Dekker


5367-4 Flynn Ch03 R2 090804

224 Chapter 3

Figure 3.114 Vapor pressure curve for ordinary liquids and helium.

Consider the gedanken experiment of sliding down the helium vapor pressure
curve, starting at the critical point, 5.2 K and 2.24 atm. As the pressure is reduced,
the vapor pressure is reduced correspondingly, reaching the normal boiling point
of 4.2 K and 1 atm. Our progress down the vapor pressure curve continues unevent-
fully until at about 2.17 K there is a totally unexpected event. A totally new liquid
phase of helium-4 is formed that has the properties of zero viscosity, anomalously
large thermal conductivity, zero absolute temperature, and zero viscosity. This is
called the superfluid.
Classical theory predicts a state of perfect order, and hence zero entropy, at
absolute zero. Liquids become solids (freeze) as the temperature is decreased, to
achieve the most orderly, crystalline state. The large zero-point energy of helium,
however, keeps the interatomic separation too large for the atoms to fall into a lat-
tice, that is, to freeze into a solid. Nevertheless, helium must still approach zero
entropy, so it does the next best thing. Instead of forming a solid under its own vapor
pressure, helium creates a second liquid phase of zero entropy, called He II. This
superfluid is helium’s answer to remaining a liquid (the atoms are too far apart
for a solid) and at the same time heading for perfect order. Helium II plays some-
thing of the role that solids do for normal fluids.
Liquid helium-4 is odorless and colorless and somewhat difficult to see in a
container because its index of refraction is so near that of the gas (nr ¼ 1.02 for liquid
He). The heat of vaporization of liquid He at the normal boiling point is 20.73 kJ=kg
(8.92 Btu=lbm), which is only 1=110 that of water.
Table 3.59 gives some of the important fixed points of helium.
Figures 3.115–3.117 are T–S charts for helium-4. The Joule–Thomson
inversion curve is given in Fig. 3.118 and data for the J-T curve are listed in
Table 3.60.

© 2005 by Marcel Dekker


5367-4 Flynn Ch03 R2 090804

Cryogenic Fluids 225

Table 3.59 Fixed Points for Helium

Critical point
T ¼ 5.2014 K (9.3625 R)
P ¼ 2.245 atm (32.99 psia; 2.275 kPa)
r ¼ 0.017399 mol=cm3 (4.348 lbm=ft3)
Normal boiling point
T ¼ 4.224 K (7.604 R)
P ¼ 1 atm (14.696 psia; 1.01325 kPa)
rgas ¼ 0.004220 mol=cm3 (1.054 lbm=ft3)
rliq ¼ 0.03122 mol=cm3 (7.802 lbm=ft3)
Lower lambda point
T ¼ 2.177 K (3.919 R)
P ¼ 0.0497 atm (0.730 psia; 5036 Pa)
rliq ¼ 0.03653 mol=cm3 (9.127 lbm=ft3
Upper lambda point
T ¼ 1.763 K (3.174 R)
P ¼ 29.74 atm (437.1 psia; 3013 M Pa
rliq ¼ 0.04507 mol=cm3 (11.26 lbm=ft3)

Source: McCarty (1972).

Helium exhibits striking properties at temperatures below 2.17 K (3.9 R). As


the liquid is cooled, instead of solidifying it changes to a new liquid phase in the
neighborhood of 2 K (3.6 R). The phase diagram of helium thus takes on an addi-
tional line separating the two phases into liquid He I at temperatures above the line
and liquid He II at lower temperatures (see Fig. 3.119). The low-temperature liquid
phase, called liquid He II, has properties exhibited by no other liquid. Helium II
expands on cooling, its conductivity for heat is enormous, and neither its heat con-
duction nor its viscosity obeys normal rules (see below). The phase transition
between the two liquid phases is identified as the lambda line, and the intersection
of the latter with the vapor pressure curve is known as the lambda point. The transi-
tion between the two forms of liquid helium, I and II, is called the lambda-point or
lambda-line because of the resemblance of the specific heat curve to the Greek letter
lambda (l) (Fig. 3.120).
Heat transfer in helium II is spectacular. When a container of liquid helium I is
pumped to reduce the pressure above the liquid, the fluid boils vigorously as the pres-
sure on the liquid decreases. During the pumping operation, the temperature of the
liquid decreases as the pressure is decreased and part of the liquid is boiled away.
When the temperature reaches the lambda point and the fluid becomes liquid helium
II, all apparent boiling suddenly stops. The liquid becomes clear and quiet, although
it is vaporizing quite rapidly at the surface. The thermal conductivity of liquid
helium II is so large that vapor bubbles do not have time to form within the body
of the fluid before the heat is quickly conducted to the surface of the liquid. Liquid
He I has a thermal conductivity of approximately 0.024 W=(m K) at 3.3 K [0.014
Btu=(hr ft  F) at 6 R], whereas liquid He II can have an apparent thermal conductiv-
ity as large as 86,500( W=(m K) [50,000 But=hr ft  F) ]—much higher than that of
pure copper at room temperature.
Not all the physical properties of helium-4 undergo such dramatic changes at
the lambda transition. For instance, there is no latent heat involved in crossing the
lambda line and no discontinuous change in volume. The lambda transition is

© 2005 by Marcel Dekker


5367-4 Flynn Ch03 R2 090804

226 Chapter 3

Figure 3.115 Temperature–entropy chart for helium-4, 2–12 K.

usually considered a transition of the ‘‘second order,’’ i.e., the Gibbs energy has a
discontinuity in the second derivative.
As mentioned earlier, He II is often referred to as a superfluid. Immediately
below the lambda point, the flow of liquid through narrow slits or channels becomes
very rapid. Figure 3.121 shows the viscosity of liquid helium measured by an oscillat-
ing disk. Helium I has a viscosity of about 3  106 (Pa s), whereas this experiment
showed that He II has a viscosity of about 1  107 Pa s at 1.3 K. However, Kapitza

© 2005 by Marcel Dekker


5367-4 Flynn Ch03 R2 090804

Cryogenic Fluids 227

Figure 3.116 Temperature–entropy chart for helium-4, 10–110 K.

and Allen and Misener showed that the viscosity for flow through thin channels
(104–105 cm) was about 1012 Pa s and was independent of the pressure drop or
length of channel and dependent only on the temperature. Thus, the viscosity of
helium below the lambda point is different for bulk flow and for flow through very
thin channels.
To explain these viscosity effects, a ‘‘two-fluid’’ model is used. The liquid is
considered to be composed of two fluids where the total density r is made up of a
normal density rn and a superfluid density component rs such that r ¼ rn þ rs.

© 2005 by Marcel Dekker


5367-4 Flynn Ch03 R2 090804

228 Chapter 3

Figure 3.117 Temperature–entropy chart for helium-4, 100–1500 K.

The former is termed the normal fluid, while the latter has been designated the
superfluid because under certain conditions the fluid acts as if it had no viscosity.
Figure 3.122 shows the temperature dependence of rn and rs below the lambda
transition. This picture is very useful in explaining many of the physical observations
of He II. However, the detailed theoretical description of the helium problem is
much more complicated, with the two-fluid model being only a first approximation.

© 2005 by Marcel Dekker


5367-4 Flynn Ch03 R2 090804

Cryogenic Fluids 229

Figure 3.118 Joule–Thomson inversion curve for helium.

To resolve the viscosity paradox, assume that at the lambda point all the fluid
is normal fluid with a normal viscosity and that at absolute zero all the fluid is super-
fluid with zero viscosity. In the thin-channel experiment described above, only the
superfluid atoms, which have zero entropy and do not interact, can flow through

Table 3.60 Joule–Thomson Inversion Curve Data for Helium

Temperature Pressure Density

K 
R atm psia mol=L lb=ft3

4.5 8.1 1.821 26.76 30.83 7.703


5 9.0 3.768 55.37 30.68 7.667
6 10.8 7.266 106.8 30.03 7.504
7 12.6 10.74 157.8 29.53 7.378
8 12.4 14.10 207.2 28.99 7.245
9 16.2 17.31 254.4 28.43 7.106
10 18.0 20.36 299.2 27.86 6.962
12 21.6 25.57 375.8 26.42 6.602
14 25.2 29.29 430.5 24.72 6.177
16 28.8 32.07 471.3 23.07 5.764
18 32.4 34.44 506.2 21.61 5.400
20 36.0 36.18 531.7 20.20 5.046
22 39.6 37.33 548.6 18.82 4.703
24 43.2 37.93 557.4 17.48 4.367
26 46.8 37.98 558.2 16.15 4.035
28 50.4 37.48 550.8 14.83 3.705
30 54.0 36.40 535.0 13.49 3.372
32 57.6 34.71 510.1 12.13 3.030
34 61.2 32.32 475.0 10.71 2.675
36 64.8 29.13 428.0 9.194 2.297
38 68.4 24.89 365.8 7.527 1.881
40 72.0 19.11 280.8 5.567 1.391
42 75.6 9.80 144.0 2.780 0.695
43 77.4 0.03 0.5 0.009 0.002

© 2005 by Marcel Dekker


5367-4 Flynn Ch03 R2 090804

230 Chapter 3

Figure 3.119 Phase diagram for helium-4.

Figure 3.120 The specific heat of saturated liquid helium.

the slit. On the other hand, the oscillating disk is damped by the normal fluid and
thus accounts for the shape of the viscosity curve below the lambda point.
The flow of He II through very thin channels is accompanied by two very inter-
esting thermal effects called the thermomechanical effect and the mechanocaloric
effect.
The thermomechanical effect was discovered in 1938 and is illustrated in
Fig. 3.123. When heat is applied to the fluid in the inner container, the superfluid
component (being cold) tends to move toward a region of higher temperature.
The superfluid can flow very rapidly through the narrow channel, whereas the
flow of the normal component is inhibited by the channel resistance. Thus, a

© 2005 by Marcel Dekker


5367-4 Flynn Ch03 R2 090804

Cryogenic Fluids 231

Figure 3.121 The total viscosity and normal viscosity of helium II.

Figure 3.122 Density vs. temperature of helium II.

thermomechanical pressure head Dh commensurate with the temperature rise DT


builds up in the inner container.
The mechanocaloric effect, Fig. 3.124, was observed in 1939. When liquid He II
is allowed to flow through a fine powder, the remaining liquid is observed to rise in
temperature. This effect can be explained by assuming that nearly all of the
fluid flowing out of the container is superfluid, thus carrying zero entropy. The
concentration of normal fluid in the liquid above the powder increases with time,
resulting in a rise in temperature.
Yet another interesting phenomenon, that of the helium film, is shown in
Fig. 3.125. A surface in contact with He II will be covered with a film of liquid. If
a beaker is lowered into a flask of He II, it will gradually fill, as shown in Fig. 3.125a.
When raised above the liquid, it will empty, as shown in Fig. 3.125b. The rate of flow
proceeds at a constant rate (at constant temperature) and is independent of the dif-
ference in level, the length of the path, and the height of the barrier over which the
flow takes place. The rate of film transfer increases with decreasing temperature,

© 2005 by Marcel Dekker


5367-4 Flynn Ch03 R2 090804

232 Chapter 3

Figure 3.123 The thermomechanical effect.

Figure 3.124 The mechanocaloric effect.

© 2005 by Marcel Dekker


5367-4 Flynn Ch03 R2 090804

Cryogenic Fluids 233

Figure 3.125 The helium film effect.

being zero at the lambda point and becoming nearly temperature-independent below
1.5 K.
Sound can be propagated in liquid He II by at least three mechanisms. First, or
ordinary, sound is the transfer of energy by a pressure wave. Second sound is a tem-
perature wave caused by out-of-phase oscillations of superfluid and normal compo-
nents of the He II. The velocity of second sound is zero at the lambda point and rises
to a value of about 20 m=s between 1 and 2 K.
Third sound is a wave motion present in helium films in which the superfluid
component oscillates but the normal component remains fixed to the walls. The wave
motion appears as an oscillation in the thickness of the film. The velocity if propaga-
tion of third sound is approximately 0.5 m=s.
Yet another form of wave propagation has been postulated to exist in helium-3
at very low temperatures; this form has been termed zero sound. The attenuation of
ordinary sound waves (first sound) in helium-3 becomes infinite as the temperature
goes to zero. In this limit, zero sound replaces first sound. Zero sound is described as
an oscillation in the
pffiffishape
ffi of the Fermi surface. In the limit of T ¼ 0, the velocity of
zero sound will be 3 times the velocity of first sound (at finite temperature). It is not
at all obvious what type of experiments should be tried to detect zero sound. The
situation is analogous to that which occurred after second sound was predicted to
exist in helium-4: Experiments to detect it failed until it was realized that thermal
rather than acoustical techniques were required.
Other Properties. Figures 3.126–3.142 give the density, thermal conductivity,
specific heats, heat of fusion, heat of vaporization, vapore pressure, surface tension,
viscosity, of helium, taken from the WADD report mentioned before (Johnson,
1960).

© 2005 by Marcel Dekker


5367-4 Flynn Ch03 R2 090804

234 Chapter 3

Figure 3.126 Density of saturated solid helium.

3.8.2. Helium-3
The isotope helium-3 is very rare and is difficult to isolate from the isotope helium-4.
Interest in helium-3 is two fold. First, its use in low-temperature refrigerators extends
the minimum attainable temperature (outside of using magnetic cooling techniques)

© 2005 by Marcel Dekker


5367-4 Flynn Ch03 R2 090804

Cryogenic Fluids 235

Figure 3.127 Density of saturated liquid helium.

to about 0.3 K. Second, its properties are of fundamental interest in relation to the-
ories of quantum statistical mechanics.
The first liquefaction of helium-3 was reported by Sydoriak, Grilly, and
Hammel in 1948. Although they had only a total of 20 cm3 of gas at STP to work

© 2005 by Marcel Dekker


5367-4 Flynn Ch03 R2 090804

236 Chapter 3

Figure 3.128 Thermal conductivity of gaseous helium near 1 atm.

with, they were able to ascertain the critical point (Tc ¼ 3.35 K, Pc ¼ 875 mmHg) and
measure the saturated vapor pessure from the critical point down to 1.2 K.
Previous to this, it had been seriously questioned whether helium-3 could be
liquefied at all under its saturated vapor pressure. The zero-point energy of
helium-3 should be four-thirds as large as that of helium-4, which is in itself very
large. It was felt that this would make the zero-point energy higher than the potential
energy associated with van der Waals attraction at all interatomic spacings and,
therefore, the gas might not liquefy at any temperature without large external pres-
sures.
The behavior of the light isotope 3He differs from that of 4He. This is partly
due to its greater zero-point energy. Since 3He has only three-fourths the mass of
4
He, its frequency of vibration is higher, and consequently the quantity (1=2) hv is
greater. This zero-point energy effectively counteracts about 95% of the cohesive
energy of 3He so that it very nearly remains a gas all the way to 0 K. Both
liquid helium-3 and liquid helium-4, in fact, display many different attributes more
characteristic of gases than of other liquid, and this is directly traceable to their high
zero-point energy. This is not the whole story, however. Helium-3 obeys a different
set of quantum statistics than helium-4 owing to its antisymmetrical nuclear struc-
ture. For our purposes, the high zero-point energy of these quantum fluids is suffi-
cient explanation.
Liquid helium-3 is a clear, colorless substance having a normal boiling point of
3.19 K (5.74 R) and a density at the normal boiling point of 58.95 kg=m3
(3.68 lbm=ft3). The heat of vaporization of liquid helium-3 at the normal boiling point
is only 8.48 kJ=kg (3.65 Btu=lbm) —so small that there was some doubt in the minds of
early investigators that helium-3 could be liquefied at normal atmospheric pressure.
As in the case of liquid helium-4, liquid helium-3 remains in the liquid state under

© 2005 by Marcel Dekker


5367-4 Flynn Ch03 R2 090804

Cryogenic Fluids 237

Figure 3.129 Thermal conductivity of liquid helium.

its own vapor pressure all the way down to absolute zero. Helium-3 must be
compressed to 28.9 bar (29.3 atm) at 0.32 K (0.576 R) before it will solidify.
An approximate pressure–temperature diagram for helium-3 is shown in Fig.
3.140 These data were used by Arp and Kropschot (1962) to prepare Fig. 3.141
From this figure the practical interest in helium-3 becomes clear. Liquid helium-3

© 2005 by Marcel Dekker


5367-4 Flynn Ch03 R2 090804

238 Chapter 3

Figure 3.130 Thermal conductivity of solid helium.

© 2005 by Marcel Dekker


5367-4 Flynn Ch03 R2 090804

Cryogenic Fluids 239

Figure 3.131 Specific heat of saturated liquid helium.

© 2005 by Marcel Dekker


5367-4 Flynn Ch03 R2 090804

240 Chapter 3

Figure 3.132 Specific heat (CV) of helium.

Figure 3.133 Specific heat (CP) of helium.

© 2005 by Marcel Dekker


5367-4 Flynn Ch03 R2 090804

Cryogenic Fluids 241

Figure 3.134 Heat of fusion of helium.

© 2005 by Marcel Dekker


5367-4 Flynn Ch03 R2 090804

242 Chapter 3

Figure 3.135 Heat of vaporization of helium.

© 2005 by Marcel Dekker


5367-4 Flynn Ch03 R2 090804

Cryogenic Fluids 243

Figure 3.136 Vapor pressure of liquid helium.

© 2005 by Marcel Dekker


5367-4 Flynn Ch03 R2 090804

244 Chapter 3

Figure 3.137 Surface tension of liquid helium.

© 2005 by Marcel Dekker


5367-4 Flynn Ch03 R2 090804

Cryogenic Fluids 245

Figure 3.138 Viscosity of liquid helium.

in equilibrium with its vapor at a given pressure is significantly colder than liquid
helium-4 at the same pressure. In an ordinary dewar system, it is difficult to reduce
the helium vapor pressure to less than a corresponding temperature of about 0.8 K.
With a modest (closed cycle) pumping system, it is not difficult to reach 0.3 K by

© 2005 by Marcel Dekker


5367-4 Flynn Ch03 R2 090804

246 Chapter 3

Figure 3.139 Viscosity of gaseous helium near 1 atm.

using helium-3. This decrease of about 0.5 K may seem small, but in reality it is more
like a gain of a factor of about 3 in temperature.
Mixtures of helium-3 and helium-4 are not completely miscible at very low
temperatures, as shown in Fig. 3.142. In fact, as absolute zero is approached,
helium-3 appears to be completely insoluble in helium-4. At other temperatures,
a phase separation occurs in any mixture whose average concentration, at the given

© 2005 by Marcel Dekker


5367-4 Flynn Ch03 R2 090804

Cryogenic Fluids 247

Figure 3.140 Pressure–temperature diagram for helium-3.

temperature, falls beneath this curve. The concentrations of the separated phases are
given by the two intersections of this curve with the (horizontal) line of constant
temperature. This separation into two liquid phases, and the difference in vapor
pressures, forms the basis for the helium-3=helium-4 dilution refrigerator used
to obtain temperatures close to absolute zero.

Figure 3.141 Vapor pressure of helium-3 and helium-4.

© 2005 by Marcel Dekker


5367-4 Flynn Ch03 R2 090804

248 Chapter 3

Figure 3.142 Phase separation in 3He–4He mixtures.

3.8.3. Properties of Helium


All helium property data are from McCarty (1972) unless otherwise noted.
a. Saturation Properties. The saturation properties of liquid helium-4 are
given in Table 3.61 for gaseous and liquid helium. Table 3.62 gives the thermal con-
ductivity of saturated liquid helium for 2.3–2.8 K.
b. Surface Tension. The surface tension for helium-4 can be calculated using
the equation
g ¼ g0 ð1  T=Tc Þ ð3:40Þ

where g0 ¼ 0.5308 dyn=cm and Tc ¼ 5.2014.


c. Heat of Vaporization. The heat of vaporization of helium is listed for a
range of temperatures in Table 3.63.
d. Dielectric Constant. The dielectric constant of a fluid may be calculated
from the Clausius–Mossotti equation
 
e  1 3M
¼P ð3:41Þ
e þ 2 4pr
where e is the dielectric constant, r is the density, and P is the specific polarizability,
a property of the substance having dimensions of specific volume. Measurements of
the dielectric constant indicated that for helium-4 the specific polarizability is a weak
function of density and that the first density correction is negative. The equation

p ¼ 0:123396  0:0014r ð3:42Þ

can be used, where p is the specific polarizability in cm3=g and r is the density in
g=cm3. The uncertainty of tabulated values of dielectric constant shown in Table 3.61
is estimated to be 0.01%.
e. Melting Curve. The melting pressure of helium can be expressed as

P ¼ 25:00 þ 0:053T 8 ð3:43Þ

© 2005 by Marcel Dekker


Table 3.61 Thermodynamic Properties of Coexisting Liquid and Gaseous Helium

Cryogenic Fluids
Spec. heat (J=(g K)) Thermal
Pressure Densitya Enthalpya Entropya conductivity Viscositya Dielectric Prandtl
Temp. (K) (atm) (g=cm3)  100 (J=g) (J=(g .K)) CV CP (mW=cm. K) (g=(cm s)  106 constant e number Npr

2.177b 0.04969 14.62


0.1177 25.41 11.92 3.20 5.61 0.0447 5.38 1.00044 0.675
2.20 0.05256 14.61 3.276 1.671 3.10 3.16 0.148 36.1 1.05500 0.778
0.1235 25.51 11.85 3.20 5.63 0.0453 5.45 1.00046 0.677
2.30 0.06629 14.58 3.567 1.796 2.46 2.56 0.155 36.7 1.05486 0.605
0.1503 25.91 11.57 3.21 5.69 0.0480 5.76 1.00056 0.683
2.40 0.08228 14.53 3.816 1.898 2.10 2.25 0.160 37.1 1.05468 0.521
0.1805 26.30 11.32 3.22 5.77 0.0508 6.07 1.00067 0.690
2.50 0.1008 14.48 4.045 1.986 1.93 2.13 0.165 37.3 1.05448 0.0482
0.2144 26.88 11.09 3.22 5.85 0.0535 6.38 1.00079 0.698
2.60 0.1219 14.42 4.269 2.068 1.85 2.10 0.168 37.3 1.05424 0.468
0.2521 27.04 10.87 3.23 5.93 0.0562 6.69 1.00093 0.707
2.70 0.1460 14.35 4.496 2.147 1.84 2.14 0.171 37.3 1.05399 0.468
0.2939 27.38 10.66 3.24 6.02 0.0589 7.01 1.00109 0.717
2.80 0.1730 14.28 4.731 2.225 1.85 2.22 0.174 37.2 1.05371 0.475
0.3401 27.71 10.47 3.25 6.13 0.0616 7.32 1.00126 0.728
2.90 0.2033 14.20 4.975 2.304 1.88 2.31 0.176 37.0 1.05341 0.487
0.3908 28.03 10.28 3.26 6.24 0.0643 7.64 1.00145 0.741
3.00 0.2371 14.11 5.231 2.382 1.92 2.42 0.178 36.8 1.05309 0.500
0.4463 28.33 10.11 3.26 6.36 0.0671 7.96 1.00165 0.755
3.10 0.2744 14.02 5.501 2.462 1.96 2.54 0.180 36.5 1.05274 0.514
0.5070 28.61 9.937 3.27 6.49 0.0699 8.29 1.00188 0.770
3.20 0.3156 13.93 5.783 2.542 2.00 2.67 0.182 36.2 1.05238 0.529
0.5731 28.87 9.774 3.28 6.64 0.0726 8.62 1.00212 0.787
3.30 0.3607 13.83 6.081 2.624 2.04 2.80 0.184 35.8 1.05199 0.545
0.6452 29.11 9.616 3.28 6.80 0.0755 8.95 1.00239 0.807

249
(Continued)

© 2005 by Marcel Dekker


250
Table 3.61 (Continued )

Spec. heat (J=(g K)) Thermal


Pressure Densitya Enthalpya Entropya conductivity Viscositya Dielectric Prandtl
Temp. (K) (atm) (g=cm3)  100 (J=g) (J=(g .K)) CV CP (mW=cm. K) (g=(cm s)  106 constant e number Npr

3.40 0.4100 13.72 6.393 2.706 2.07 2.95 0.186 35.5 1.05157 0.562
0.7235 29.32 9.463 3.29 6.98 0.0784 9.30 1.00268 0.828
3.50 0.4637 13.60 6.722 2.790 2.11 3.10 0.188 35.1 1.05113 0.580
0.8085 29.52 9.314 3.30 7.18 0.0813 9.65 1.00300 0.852
3.60 0.5220 13.48 7.068 2.875 2.14 3.28 0.190 34.7 1.05066 0.599
0.9008 29.69 9.169 3.30 7.40 0.0843 10.0 1.00334 0.878
3.70 0.5849 13.35 7.432 2.962 2.18 3.47 0.191 34.2 1.05016 0.621
1.001 29.84 9.025 3.31 7.66 0.0874 10.4 1.00371 0.908
3.80 0.6528 13.21 7.816 3.050 2.21 3.68 0.192 33.8 1.04963 0.646
1.109 29.96 8.884 3.32 7.94 0.0906 10.7 1.00411 0.942
3.90 0.7257 13.06 8.221 3.141 2.25 3.92 0.194 33.3 1.04906 0.674
1.228 30.06 8.743 3.33 8.27 0.0939 11.1 1.00455 0.980
4.00 0.8040 12.90 8.650 3.234 2.28 4.19 0.195 32.8 1.04845 0.707
1.356 30.12 8.603 3.33 8.65 0.0974 11.5 1.00503 1.02
4.10 0.8878 12.73 9.105 3.330 2.32 4.51 0.196 32.3 1.04780 0.745
1.496 30.15 8.463 3.34 9.10 0.101 11.9 1.00555 1.07
4.20 0.9772 12.54 9.508 3.429 2.36 4.88 0.196 31.8 1.04710 0.791
1.649 30.14 8.322 3.35 9.63 0.105 12.4 1.00612 1.13
4.224 1.000 12.50 9.711 3.454 2.37 4.98 0.196 31.7 1.04692 0.884
1.689 30.13 8.287 3.35 9.78 0.106 12.5 1.00626 1.15
4.30 1.073 12.35 10.10 3.532 2.40 5.32 0.197 31.3 1.04635 0.847
1.818 30.08 8.177 3.35 10.3 0.110 12.8 1.00674 1.20
4.40 1.174 12.13 10.66 3.640 2.44 5.86 0.197 30.7 1.04553 0.916
2.005 29.98 8.029 3.36 11.1 0.114 13.3 1.00744 1.28

Chapter 3
4.50 1.282 11.89 11.25 3.753 2.49 6.55 0.199 30.1 1.04463 0.994
2.213 29.81 7.876 3.37 12.1 0.128 13.8 1.00821 1.39
4.60 1.397 11.63 11.90 3.873 2.54 7.44 0.200 29.5 1.04364 1.10

© 2005 by Marcel Dekker


Cryogenic Fluids
2.449 29.58 7.714 3.37 13.5 0.127 14.3 1.00909 1.52
4.70 1.519 11.34 12.61 4.002 2.59 8.68 0.202 28.9 1.04252 1.24
2.179 29.25 7.540 3.38 15.5 0.135 14.8 1.01010 1.70
4.80 1.648 11.01 13.40 4.144 2.66 10.5 0.205 28.2 1.04125 1.45
3.037 28.80 7.350 3.38 18.5 0.147 15.5 1.01128 1.95
4.90 1.784 10.61 14.30 4.304 2.73 13.6 0.210 27.4 1.03974 1.78
3.425 28.18 7.133 3.39 23.6 0.164 16.2 1.01273 2.32
5.00 1.929 10.11 15.39 4.495 2.81 19.9 0.222 26.4 1.03785 2.37
3.930 27.28 6.871 3.39 34.6 0.198 17.0 1.01461 2.97
5.10 2.082 9.489 16.82 4.747 2.92 38.5 0.263 25.1 1.03520 3.68
4.680 25.83 6.513 3.37 71.5 0.301 18.1 1.01742 4.30
5.201b 2.245 6.964 21.36 5.589
6.964 21.36 5.589
a
Top line of data ¼ saturated liquid properties; bottom line data ¼ saturated vapor properties.
b
Lambda point and critical point
Source: McCarty (1972).

251
© 2005 by Marcel Dekker
5367-4 Flynn Ch03 R2 090804

252 Chapter 3

Table 3.62 Thermal Conductivity of Saturated Liquid Helium

Thermal conductivity Thermal conductivity


Temp. (K) (mW=(cm K)) Temp. (K) (mW=(cm K))

2.3 0.181 3.0 0.214


2.4 0.185 3.5 0.238
2.6 0.195 4.0 0.262
2.8 0.205 4.2 0.271

Source: Grenier (1951).


Note: Thermal conductivity is a linear function of temperature between 2.5 and 4.5 K.

Table 3.63 Heat of Vaporization of Helium

Temp. (K) hv(J=g) Temp. (K) hv(J=g)

2.20 22.8 4.00 21.9


2.40 23.1 4.20 20.9
2.60 23.3 4.40 19.7
2.80 23.5 4.60 18.0
3.00 23.7 4.80 15.6
3.20 23.6 5.00 12.0
3.40 23.5 5.10 8.99
3.60 23.2 5.15 6.70
3.80 22.7 5.18 4.00

Source: Berman and Mate (1958).

from 1.0 to 1.4 K and as


 1:554
P T
¼ 1 ð3:44Þ
16:45 0:992

above 4 K, where P is expressed in atmospheres.


f. Vapor Pressure. The saturated vapor curve of helium is used as the interna-
tional temperature scale from 0.5 to 5.2 K. The 1958 international scale is repro-
duced in part in Table 3.64.
g. Specific Heat. The specific heat of helium at constant pressure (CP) and at
constant volume (CV) is listed in Tables 3.65 and 3.66, respectively.
Symbols:, rsat ¼ saturation pressure or vapor pressure, r ¼ density, CP ¼ speci-
specific heat at constant pressure, m ¼ viscosity, k ¼ thermal conductivity, hfg ¼ heat
of vaporization, Npr ¼ mCP=k ¼ Prandtl number, sL ¼ surface tension, e ¼ dielectric
constant.

3.8.4. Uses of Helium


Except for hydrogen, helium is the most abundant element found throughout the
universe. The helium content of the earth’s atmosphere is about 1 part in 200,000.
While it is present in various radioactive minerals as a decay product, the bulk of
the Western world’s supply is obtained from wells in Texas, Oklahoma, and Kansas.

© 2005 by Marcel Dekker


Table 3.64 Vapor Pressure of Helium-4 P in 10-3 mmHg (Hg at 0 C) g ¼ 980.665 cm=sec2

Cryogenic Fluids
T (K) 0.00 0.01 0.02 0.03 0.04 0.05 0.06 0.07 0.08 0.09

0.5 0.016342 0.022745 0.031287 0.042561 0.057292 0.076356 0.10081 0.13190 0.17112 0.22021
0.28121 0.35649 0.44877 0.56118 0.69729 0.86116 1.0574 1.2911 1.5682 1.8949
2.2787 2.7272 3.2494 3.8549 4.5543 5.3591 6.2820 7.3365 8.5376 9.9013
11.445 13.187 15.147 17.348 19.811 22.561 25.624 29.027 32.800 36.974
41.581 46.656 52.234 58.355 85.059 72.386 80.382 89.093 98.567 108.853
1.0 120.000 132.070 145.116 159.198 174.375 190.711 208.274 227.132 247.350 269.006
292.169 316.923 343.341 371.512 401.514 433.437 467.365 503.396 541.617 582.129
625.025 670.411 718.386 769.057 822.527 878.916 938.330 1000.87 1066.67 1135.85
1208.51 1284.81 1364.83 1448.73 1536.61 1628.62 1724.91 1825.58 1930.79 2040.67
2155.35 2274.99 2399.73 2529.72 2665.08 2805.99 2952.60 3105.04 3263.48 3428.07
1.5 3598.97 3776.32 3960.32 4151.07 4348.79 4553.58 4765.68 4985.18 5212.26 5447.11
5689.88 5940.76 6199.90 6467.42 6743.57 7028.47 7322.31 7625.21 7937.40 8259.02
8590.22 8931.18 9282.06 9643.02 10014.3 10395.9 10788.2 11191.2 11605.1 12030.1
12466.1 12913.7 13372.8 13843.6 14326.1 14820.7 15327.3 15846.3 16377.7 16921.7
17478.2 18047.7 18630.1 19225.5 19834.1 20455.9 21091.1 21739.7 22402.0 23077.9
2.0 23767.4 24470.9 25188.1 25919.2 26664.2 27423.3 28196.3 28983.2 29784.2 30599.1
31428.1 32271.1 33128.0 33998.6 34882.8 35780.3 36690.9 37614.3 38550.2 39500.3
40465.6 41446.6 42443.5 43456.5 44485.7 45531.3 46593.5 47672.5 48768.6 49881.8
51012.3 52160.2 53325.8 54509.2 55710.5 56930.0 58167.8 59423.8 60698.8 61992.0
63304.3 64635.2 65985.4 67354.8 68743.5 70152.0 71580.2 73028.1 74496.0 75984.2
2.5 77493.1 79022.2 80572.2 82142.9 83734.6 85347.2 86981.2 88636.7 90313.8 92012.6
93733.1 95476.0 97240.8 99028.2 100838 102669 104525 106403 108304 110228
112175 114145 116139 118156 120198 122263 124353 126465 128603 130765
132952 135164 137401 139663 141949 144260 146597 148961 151349 153763
156204 158671 161164 163684 166230 168802 171402 174028 176682 179364
3.0 182073 184810 187574 190366 193187 196037 198914 201820 204755 207719
210711 213732 216783 219864 222975 226115 229285 232484 235714 238974

(Continued)

253
© 2005 by Marcel Dekker
254
Table 3.64 (Continued )

T (K) 0.00 0.01 0.02 0.03 0.04 0.05 0.06 0.07 0.08 0.09

242266 245587 248939 252322 255736 259182 262658 266166 269706 273278

(Continued)
276880 280516 284183 287883 291615 295380 299178 303008 306871 310768
314697 318659 322654 326684 330747 334845 338976 343141 374341 351575
3.5 355844 360147 364485 368860 373269 377714 382194 386710 391262 395849
400471 405130 409825 414556 419324 424128 428968 433846 438760 443713
448702 453729 458794 463897 469038 474218 479435 484691 489985 495317
500688 506098 511547 517036 522564 528132 533739 539387 545075 550805

5367-4 Flynn Ch03 R2 090804


556574 562383 568234 574126 580059 586034 592051 598110 604210 610352
4.0 616537 622764 629033 635345 641700 648099 654541 661026 667554 674125
680740 687399 694103 700851 707643 714479 721360 728285 735255 742269
749328 756431 763579 770772 778010 785294 792623 799999 807422 814893
822411 829978 837592 845255 852966 860725 868533 876390 884296 892252
900258 908313 916418 924573 932778 941033 949338 957693 966099 974556
4.5 983066 991628 1000239 1008905 1017621 1026390 1035213 1044087 1053014 1061995
1071029 1080144 1089254 1098449 1107699 1117002 1126359 1135772 1145239 1154761
1164339 1173972 1183662 1193407 1203209 1213066 1222981 1232955 1242983 1253069
1263212 1273414 1283673 1293991 1304367 1314802 1325297 1335850 1346462 1357136
1367870 1378662 1389516 1400429 1411404 1422438 1433533 1444690 1455911 1467191
5.0 1478535 1489940 1501409 1512940 1524535 1536192 1547912 1559698 1571546 1583458
1595437 1607481 1619589 1631761 1644000 1656305 1668673 1681108 1693612 1706180
1718817 1731521 1744290

Source: Brickwedde et al. (1960).

Chapter 3
© 2005 by Marcel Dekker
5367-4 Flynn Ch03 R2 090804

Cryogenic Fluids 255

Table 3.65 Specific Heat (CP) of Helium (cal=g K)

Temp. (K) 3 atm 5 atm 6 atm 10 atm 15 atm 30 atm 50 atm 70 atm

6 2.91 1.18 0.97 0.77 0.67 0.615


6.5 2.14 3.40 2.25 1.35 1.09
7 1.84 2.83 2.84 1.53 1.20 0.90 0.74 0.71
8 1.55 1.96 2.17 1.93 1.41 1.02 0.87 0.79
9 1.46 1.67 1.78 1.96 1.60 1.14 0.91 0.87
10 1.40 1.53 1.61 1.81 1.71 1.25 1.05 0.95
12 1.42 1.46 1.59 1.64 1.40 1.18 1.08
14 1.37 1.39 1.47 1.52 1.47 1.28 1.18
16 1.34 1.35 1.40 1.45 1.48 1.37 1.22
18 1.32 1.33 1.36 1.41 1.46 1.41 1.32
20 1.31 1.34 1.38 1.44 1.43 1.37

Source: Lounasmaa (1958).

The only helium plant in the western world outside the United States is near Swift
Current, Saskatchewan.
Helium is a low-density inert gas with a combination of unique properties that
make its availability essential for a wide range of industrial and scientific endeavors.
These include applications not only in cryogenics, but also in special welding
technology, space exploration, chromatography, heat transfer, and controlled
atmosphere.
a. Pressurizing and Purging. Most of the helium for pressurizing and purging
is used in the space program. Helium is used to pressurize liquid oxygen, liquid
hydrogen, kerosene, and the hypergolic propellants. Helium is used also to pressurize
pneumatic systems on boosters and spacecraft. A Saturn booster such as that used
on the Apollo lunar missions requires about 13  106 ft3 of helium for a firing plus
more for checkouts. At the height of the Apollo program, almost one-third of all
domestic helium production was used for this purpose.

Table 3.66 Specific Heat (CV) of Helium (cal=(g K))

Temp. (K) 3 atm 5 atm 10 atm 20 atm 40 atm 60 atm 80 atm 100 atm

3 0.444 0.408 0.365 0.318


4 0.555 0.534 0.504 0.464 0.415 0.382 0.358 0.345
5 0.621 0.604 0.575 0.538 0.492 0.459 0.434 0.414
6 0.746 0.664 0.626 0.593 0.552 0.523 0.501 0.480
6.5 0.751
7 0.737 0.719 0.669 0.633 0.600 0.574 .0555
8 0.738 0.730 0.699 0.664 0.634 0.613 0.596
10 0.739 0.736 0.722 0.701 0.680 0.667 0.654
12 0.740 0.740 0.732 0.721 0.712 0.700 0.796
16 0.741 0.742 0.740 0.736 0.730 0.730
20 0.750 0.752 0.755 0.758 0.760

Source: Lounasmaa (1958).

© 2005 by Marcel Dekker


5367-4 Flynn Ch03 R2 090804

256 Chapter 3

b. Controlled Atmospheres. Helium is used to maintain a controlled atmo-


sphere for cooling vacuum furnaces in processing fuel elements for nuclear reactors
and for producing germanium and crystal growth.
c. Research. Uses include aerodynamic research in wind tunnels and shock
tubes to stimulate velocities up to about 50 times the speed of sound, the develop-
ment of improved seals and valves for positive shutoff and leak proof operation,
pharmaceutical and biological research, particle investigations in physics, develop-
ment of improved light sources, radiation detection devices, plasma arc studies, mass
transfer studies, tensile and impact tests of materials, solid-state physics, and the
development of new analytical techniques.
d. Welding. Helium alone or mixed with other gases, usually argon, is used as
a shield in the high-speed mass production welding of tubes from strip material; ship,
aircraft, spacecraft, and rocket structures; food handling equipment; diesel engine
parts; hardware; electrical devices; storage tanks; vessels and piping; and other pro-
ducts made from stainless steel, 9% nickel steel, aluminum, copper, titanium, zirco-
nium, and other metals.
e. Lifting Gas. Until after World War II, the principal use of helium was for
the inflation of military lighter-than-air craft. Although such craft are no longer in
operation, helium continues to be used as a lifting gas because it is much safer than
hydrogen. Although its density is almost twice that of hydrogen, it has about 98% of
the lifting power of hydrogen. At sea level, 28 m3 (1000 ft3) of helium lifts 31 kg
(68.5 lbm) while the same volume of hydrogen lifts 35 kg (76 lbm). The Weather
Bureau uses helium in about half of its weather balloon operations. Commercial uses
of helium include its use in blimps for advertising purposes and in special balloons
used to remove logs from mountains in inaccessible locations.

© 2005 by Marcel Dekker


5367-4 Flynn Ch04 R2 090904

4
Mechanical Properties of Solids

1. INTRODUCTION

A knowledge of the properties and behavior of materials used in any cryogenic sys-
tem is essential for proper design considerations. Often the choice of materials for
the construction of cryogenic equipment will be dictated by consideration of
mechanical and physical properties such as thermal conductivity (heat transfer along
a structural member), thermal expansivity (expansion and contraction during cycling
between ambient and low temperatures), and density (weight of the system relative to
its volume). Since properties at low temperatures are often significantly different
from those at ambient temperature, there is no substitute for test data, and fortu-
nately there are now several excellent data compilations (McClintock and Gibbons,
1960; Durham et al., 1962; Campbell, 1974; Johnson, 1960). To help make sense out
of all of the data that do exist, and to help estimate properties when no data are
present, it is useful to have certain general rules in mind. Such is the purpose of
the following discussion.
In any branch of engineering design, the choice of material is dictated by ques-
tions of safety and economy. The range of choice in cryogenic design is perhaps lim-
ited by the issue of low-temperature embrittlement. The purpose of this section is to
show which materials that are ductile at ordinary temperatures retain their ductility
at low temperatures. This determination requires an explanation of why ductility
exists at all in some materials and not others and a knowledge of the temperature-
related mechanisms that govern ductility. Common indices of strength of materials
are also explored along with the temperature dependence of this important property.

2. STRENGTH, DUCTILITY, AND ELASTIC MODULUS

When a bar of a structural solid is subjected to an elongating force, the first response
is a slight stretching that is directly proportional to the force producing it. If the
elongating force is released during this initial period, the material will return to very
nearly its original length. This is called elastic behavior, and all solid materials exhi-
bit it in some measure. At some higher value of stress, however, the material will no
longer behave elastically. It may break without further deformation (brittle beha-
vior), or it may take on a permanent deformation (ductile behavior).
The ductile or brittle behavior of structural materials is usually determined by
the familiar stress–strain test shown in Fig. 4.1. The figure represents a steel bar of
cross-section a, on which a gauge length l has been marked. Let P be the value of any
257

© 2005 by Marcel Dekker


5367-4 Flynn Ch04 R2 090904

258 Chapter 4

Figure 4.1 Tensile test of a structural material.

axial tensile load that gradually increases from zero value until the steel bar breaks.
As the value of P increases, values of the elongation e are taken.
Stress:
S ¼ P=A
Strain:
E ¼ e=l
The relationship between s and E is the stress–strain curve found experimentally and
is shown in Fig. 4.2.
Another class of materials (nonmetals) are capable of extreme great elastic
deformations. These are called elastomeric materials.
Usually, materials exhibiting plastic deformation under stress are the more
desirable for structures. Ductility is desirable so that accidental stresses beyond
design values can be redistributed to safer levels by means of plastic flow.
Brittle materials have no such mechanism to protect them from excessive stress.
When a local overstress occurs in a brittle material, the result is often failure of the
piece rather than a deformation. It is useful to know, therefore, which structural
materials that are acceptable (ductile) at normal temperatures will remain ductile
at low temperatures. This requires that we know the mechanisms of ductility in solids
and how those mechanisms change with temperature. To begin, let us examine the
structure of solids.

Figure 4.2 The stress–strain relationship of a ductile material (a) and a brittle material (b).

© 2005 by Marcel Dekker


5367-4 Flynn Ch04 R2 090904

Mechanical Properties of Solids 259

3. THE STRUCTURE OF SOLIDS

The theory of elasticity treats solids as continuous elastic media. Solids are not
continuous media. They are composed of atoms bound together in more or less
regular arrays. Essentially, three different types of solid structures exist, represented
by glasses, plastics, and metals.

3.1. Glasses
The structure of an idealized glass is shown in Fig. 4.3 (Jones, 1956). The chemi-
cal composition of this glass is G2O3, where G represents a metallic element and
O represents oxygen. The principal characteristic of the glass to be noted is the lack
of any spatial order, even though the stoichiometric relationship between G and O is
observed. Liquids also show such lack of order, and glasses can be described to a first
approximation as extremely viscous liquids.

3.2. Plastics
Plastics, or polymers, are composed of giant long-chain molecules. Polymer mole-
cules may have from tens to thousands of atoms each and are essentially linear. They
lie in tangled disarray; a segment of a typical polymer molecule is shown in Fig. 4.4.
The intermolecular force that binds the polymer molecules to one another is the
rather weak van der Waals force. Some cross-linking among the polymer molecules
may also be present. Cross-linking is actual chemical bonding between adjacent
molecules and is much stronger than the van der Waals forces. The more highly
cross-linked (thermoset) polymers form much more rigid solids than those that are
only sparingly cross-linked (thermoplastics).

3.3. Metals
Compared to glasses and polymers, metals have a highly ordered structure. Metal
atoms are arranged in symmetrical crystal lattices. The three most common metal
crystal lattices are shown in Fig. 4.5 (Quarrel, 1959). The face-centered cubic (fcc)

Figure 4.3 Idealized structure of glass.

© 2005 by Marcel Dekker


5367-4 Flynn Ch04 R2 090904

260 Chapter 4

Figure 4.4 Structure of a representative polymer.

lattice consists of a cube that has an atom at each of its eight corners and one at the
center of each of the six crystal faces. The hexagonal close-packed (hcp) lattice con-
sists of a right hexagonal prism with an atom at each of its 12 vertices, an atom at the
center of each of its two regular hexagonal ends, and three more atoms located mid-
way between the ends of the prism, each on a line parallel with the vertical axis of the
prism and running through the center of the equilateral triangle formed by three
neighboring atoms in the end faces. The body-centered cubic (bcc) lattice is a cube
with an atom at each of its eight corners and an atom at the center of the body of
the cube. Other more complicated metal structures exist, but only these three basic
types will be considered here.
The drawings of Fig. 4.5 are useful for visualizing the slip planes that exist in
various metal crystal systems. For example, from the fcc lattice shown in Fig. 4.5a, it
is easy to see how planes of atoms might slide rather easily over one another from
upper left to lower right. This lattice also has several other such slip planes. Although
it is not readily apparent from the figure, a three-dimensional model of these crystal
systems would show that the bcc lattice offers the least number of slip planes of the
three shown, and the hcp system falls in between the fcc and bcc systems.
Real crystals do not have the perfect geometric arrangement of atoms shown in
Fig. 4.5 (Van Buren, 1960). For example, there are always some impurity atoms
present. These impurity atoms cause irregularities in the crystal lattice. Real
crystals may also have locations where atoms are simply out of place. An extra atom
in the interstices between regular crystal sites is called an interstitial atom. A
missing atom is called a vacancy. Interstitial atoms and vacancies are examples of
‘‘point defects’’ in the crystal.
Extended imperfections can also occur. An extra plane of atoms can be
present in a crystal lattice. Extended imperfections of this type are called edge
dislocations.
Another type of extended imperfection is common in crystals: the screw dislo-
cation. This type of dislocation occurs when part of a crystal has slipped one atomic
distance relative to its adjacent part.

Figure 4.5 The three most common crystal structures of metals.

© 2005 by Marcel Dekker


5367-4 Flynn Ch04 R2 090904

Mechanical Properties of Solids 261

Both dislocations and point defects are nearly always present to some degree in
real metallic crystals (Cottrell, 1953). The concept of a geometrically perfect crystal is
useful in explaining certain properties of crystals, but there are other properties that
can be adequately explained only by consideration of the defects. The concept of
dislocations, for example, helps explain the yield stress and ductility of metals.
The regular arrangement of atoms in the crystal (and not so much its defects) helps
explain the modulus of elasticity.

4. DUCTILITY

As mentioned, the role of dislocations is useful in explaining the yield strength and
ductility of certain crystalline solids.
One could account for elasticity by imagining a regular crystal structure with-
out dislocations such as those of Fig. 4.5. If a certain horizontal shearing stress were
applied to the top of the crystal lattice, then the structure would bend, but it would
snap back to its original shape when the stress was removed. Above a certain stress,
however, plastic shear would suddenly take place along one of the planes of atoms,
resulting in a permanent deformation.
The ‘‘perfect crystal’’ was the model originally used to explain the plastic defor-
mation of metals. However, when the stress necessary to shear a perfect crystal was
calculated on the basis of interatomic forces, the results were a thousand times
greater than the experimentally observed stresses. Taking the presence of disloca-
tions into account gives better results, because the stress required to force a disloca-
tion through a crystal is of the required magnitude (Barrett, 1957). The movement of
one plane of atoms over another in a crystal actually occurs progressively as the
dislocation moves through the crystal. The resulting crystal deformation is the same
as if the process had taken place suddenly. Shearing a perfect crystal lattice structure
is like moving a heavy rug all at once; forcing a dislocation through the crystal is
like moving a wrinkle through the rug. The displacement of the rug is the same in
both cases.
The ability of crystalline materials to deform plastically (their ductility)
depends primarily on the mobility of dislocations within the crystal (Fisher, 1957).
The mobility of dislocations depends on temperature, as one might expect, and on
several other things as well. The number of slip systems available in a crystal influ-
ences dislocation mobility. Slip systems are directions within the crystal in which the
planes can slip easily over one another. Dislocations can move most easily through a
crystal in the direction of slip planes.
The movement of dislocations can also be impeded in many ways. Impurity
atoms at lattice sites can lock dislocations in place. Vacancies and interstitial atoms
also provide obstacles to the movement of dislocations. The existence of hard foreign
particles in the crystal can also impede the mobility of dislocations.
Grain boundaries, which are discontinuities in the orientation of neighboring
crystals, also are obstacles. A metal having a small grain size will have a higher yield
stress than one of large grain size. One reason annealed metals are more ductile than
work-hardened ones is that work hardening reduces the size of grains. Finally,
dislocations themselves provide obstacles to the progress of other dislocations.
Glasses have no slip systems and are therefore very brittle. When stress is
applied to a piece of glass, the atomic bonds rupture at some point on the surface.
The crack then propagates, causing a fracture. The final stage of rupture of a metal

© 2005 by Marcel Dekker


5367-4 Flynn Ch04 R2 090904

262 Chapter 4

is glass-like. After plastic deformation, dislocation motion is blocked by tangles of


the great number of dislocations produced by work hardening. Greater stress then
ruptures atomic bonds, much like a glass, at some location where the stress is too
great for the deformed crystal. Further stress then propagates the crack through
the material (fracture).
Plastics and elastomers that are not highly cross-linked can yield considerably
to tension stress by merely uncoiling their long molecules. As only van der Waals
bonds are ruptured in this process, elongation takes place easily. The long molecules
merely slide over each other. When this sliding process is completed, the force of the
chemical bonds comes into play.
The resistance of chemical bonds in cross-linked plastics or elastomers comes
into play early, and these materials elongate less under stress. Highly cross-linked
plastics, in fact, are so brittle that they resemble glass.

4.1. Low-Temperature Ductility


The effect of temperature on ductility is, to a first approximation, the effect of
temperature on the response mechanisms discussed above—the crystal system
(slip planes) and dislocation motion. This will be discussed first for metals, then
for elastomers. Embrittlement of these materials at low temperatures is summarized
in Table 4.1.

4.1.1. Metals
Since the number of slip systems is not usually a function of temperature, the ducti-
lity of fcc metals is relatively insensitive to a decrease in temperature (Brick, 1953).
Metals of other crystal lattice types tend to become brittle at low temperatures. Crys-
tal structure and ductility are related because the fcc lattice has more slip systems
than the other crystal structures. In addition, the slip planes of bcc and hcp crystals
tend to change at low temperature, which is not the case for fcc metals. Therefore,
copper, nickel, all of the copper–nickel alloys, aluminum and its alloys, and the

Table 4.1 Embrittlement of Structural Materials at Low Temperatures

Materials that remain ductile at low


temperatures, if ductile at room Materials that become brittle at
temperature low temperatures

Copper Iron
Nickel Carbon and low alloy steels
All copper–nickel alloys
Aluminum and all its alloys Molybdenum
Austenitic stainless steels containing Niobium
more than 7% nickel Zinc
Most bcc metals
Zirconium Most plastics
Titanium
Most fcc metals
Polytetrafluoroethylene

© 2005 by Marcel Dekker


5367-4 Flynn Ch04 R2 090904

Mechanical Properties of Solids 263

austenitic stainless steels that contain more than approximately 7% nickel, which are
all face-centered cubic, remain ductile down to low temperatures if they are ductile at
room temperature. Iron, carbon and low-alloy steels, molybdenum, and niobium,
which are all body-centered cubic, become brittle at low temperatures. The hexago-
nal close-packed metals occupy an intermediate place between fcc and bcc behavior.
Zinc undergoes a transition to brittle behavior in tension, whereas zirconium and
pure titanium remain ductile (Corruccini, 1957).
The thermal vibration of atoms in the crystal lattice is strongly temperature-
dependent and is less effective in assisting dislocation motion at low temperatures.
The interaction of dislocations with thermal vibrations is complicated, but it is non-
etheless satisfying to find that ductility usually decreases somewhat with a decrease
in temperature.
The complete situation regarding brittleness in metals at low temperature
depends on more than just crystal structure. For instance, bcc potassium and beta
brass remain ductile down to 4.2 K, and the bcc metals lithium and sodium show
no signs of brittleness down to 4.2 K. In general, however, brittleness will not
occur in face-centered cubic metals in which dislocations cannot be firmly locked by
impurity atoms.

4.1.2. Elastomers
As mentioned, the tendency to show a ductile–brittle transition is correlated with the
lattice type. Thus, the fcc metals show only a few cases of this effect and for struc-
tural purposes may be regarded as almost uniformly well behaved. These include
copper, nickel, aluminum, the solid solution alloys of each of these, and the austeni-
tic stainless steels. In contrast, bcc metals for the most part show brittle behavior
(though the transition zones of some can be depressed to low temperatures). The fer-
ritic steels are by far the most prominent of these. Prominent among the established
structural metals with a different lattice are the hexagonal close-packed metals mag-
nesium and titanium. The impact strength of magnesium is low at all subambient
temperatures because its brittle transition zone is above room temperature. Tests
on commercial titaniums indicate that ductility is retained in tension to low tempera-
tures if the amounts of the interstitial solutes carbon, oxygen, nitrogen, and hydro-
gen are small. However, notched-bar impact tests show a transition above ambient
temperature.
For temperatures much below 200 K, it is common practice to use the
face-centered cubic metals almost exclusively, especially where shock and vibration
are encountered. However, less expensive steels can be used in many less critical
applications, especially for temperatures above 150–200 K.
All but one plastic or elastomer become brittle at low temperatures. Polytetra-
fluoroethylene (PTFE) is unique in that it can still be deformed plastically to a small
degree at 4 K.
Plastics and elastomers do not respond to stress as metals do. The less cross-
linked elastomers yield by uncoiling their long-chain molecules and by sliding over
one another. The thermal energy of the material at room temperature facilitates this
motion.
At low temperatures, however, the attractive intermolecular forces are more
effective than the thermal energy ‘‘lubricant,’’ and the material deforms less readily.
This effect is especially pronounced through its ‘‘glass transition’’ in the temperature
range at the onset of brittleness.

© 2005 by Marcel Dekker


5367-4 Flynn Ch04 R2 090904

264 Chapter 4

4.2. Ductility as a Function of Temperature


The curve of ductility (elongation in a tensile test, for example) vs. temperature gen-
erally has an S shape similar to that shown in Fig. 4.6. The material shown in Fig. 4.6
has considerable ductility at higher temperatures but is brittle at low temperatures.
All materials have a ductility temperature dependence represented by at least some
portion of Fig. 4.6. Between temperatures T1 and T2, there is a transition from duc-
tile to brittle behavior. This brittle transition may occur in any temperature range,
wide or narrow, or, for some materials, may not occur at all. If the material is either
ductile or brittle over the entire temperature range, then the curve consists only of
the portion above T2 or below T1. For example, face-centered cubic metals show
only the curve above T2, and brittle materials such as glass show only the curve
below T2.
Some hard copper alloys appear to undergo a gradual transition over most of
the temperature range, and a similar graph for these materials looks like a portion of
the region between T1 and T2 but spread out over a considerably greater temperature
range.
Ductility is well represented by the entire curve in Fig. 4.6 for steels except for
the fcc austenitic types. These austenitic steels display ductility near room tempera-
ture but undergo a transition to brittle behavior at some temperature below room
temperature.
The brittle transition region is that range of temperatures where the material’s
important mechanisms of response to stress are becoming inactive. These response
mechanisms include the mobility of dislocations for a metal and the ability of the
molecules to slide over one another in a polymer.
A particular case of brittle behavior is shown by tensile data on a low-carbon
steel in Fig. 4.7 (Corruccini, 1957). Two features are characteristic of these materials:
1. The large decrease in elongation and reduction of area that occur in a rela-
tively narrow region of temperature, the brittle transition zone
2. The rapid rise in yield strength, which approaches the material’s tensile
strength as the temperature is lowered through the transition zone
In contrast, Fig. 4.8 shows a similar set of properties (Corruccini, 1957)
for Type 347 stainless steel, an alloy that is apparently not brittle at any low
temperature. The two main characteristics of these alloys are (1) the relative
constancy of the yield strength and the increased capacity for work hardening at
lower temperatures as shown by the steep rise of the ultimate strength, and (2) the

Figure 4.6 The general curve of ductility vs. temperature.

© 2005 by Marcel Dekker


5367-4 Flynn Ch04 R2 090904

Mechanical Properties of Solids 265

Figure 4.7 Brittle behavior of a low-carbon steel.

maintenance of ductility at all temperatures as shown by the high values of


elongation and reduction of area.
The curves of ductility vs. temperature for several important structural materi-
als are shown in Fig. 4.9 (Durham et al., 1962).

4.3. Ductility as a Function of Strain Rate and Stress Complexity


Increasing strain rate and increasing complexity of the stress system have the effect
of decreasing ductility. Two tests of mechanical properties, the tensile and impact
tests, are used to measure the effect of strain rate and stress complexity on ductility.
The tensile test involves unidirectional stresses applied at comparatively slow rates.
The impact test applies stresses in several directions at rapid rates. At a given

Figure 4.8 Ductile behavior of a Type 347 stainless steel.

© 2005 by Marcel Dekker


5367-4 Flynn Ch04 R2 090904

266 Chapter 4

Figure 4.9 The curve of ductility vs. temperature for several important structural materials.
(1) 2024 T4 aluminum; (2) beryllium copper; (3) K Monel; (4) titanium; (5) 304 stainless steel;
(6) C1020 carbon steel; (7) 9% nickel steel.

temperature, a material may exhibit considerable ductility in the tensile test but not
in the impact test.
Figure 4.10 shows this effect for an ordinary carbon steel. This figure illustrates
that the ductility of a material is affected by the type of stress system and the rate of
application of this stress system. Between T2 and T3, for example, the carbon steel
displays ductile behavior in a simple uniaxial stress system (tensile test) and displays
brittle characteristics at high rates of loading (impact test). Increasing either the
strain rate or the complexity of the stress system moves the curves to the right. This
amounts to an increase in the brittle transition temperature. Uniformly ductile or
brittle behavior is observed above T4 and below T1.
The tensile elongation and the impact energy vs. temperature of annealed oxy-
gen-free copper are shown in Fig. 4.11. The tensile elongation of copper is quite high

Figure 4.10 Effect of strain rate and stress complexity.

© 2005 by Marcel Dekker


5367-4 Flynn Ch04 R2 090904

Mechanical Properties of Solids 267

Figure 4.11 Tensile elongation and impact energy of annealed oxygen-free copper.

at room temperature, and it even increases somewhat at low temperatures. The


impact energy curve follows the same trend. Therefore, copper will exhibit no brittle
behavior at low temperatures for the usual applications. All of the common copper
alloys investigated so far are ductile at low temperatures if they are ductile at higher
temperatures (Teed, 1950). Copper alloys that have a predominantly face-centered
cubic crystal structure have a ductility largely independent of temperature.
The behavior of another fcc alloy, one of the chromium–nickel stainless steels,
is shown in Fig. 4.12. There is a small decrease in tensile elongation and impact
energy with decreasing temperature. The ductility of this group of alloys is largely
independent of temperature. When the nickel is removed from these stainless steel
alloys, they crystallize into the body-centered cubic lattice. Type 430 stainless steel
is typical of this group, and its low-temperature behavior is shown in Fig. 4.13. At
about 200 K, this alloy has become brittle in the impact test. Below about 75 K, it
displays very little ductility even in the tensile test.
These transitions can become very sharp, as shown in Fig. 4.14. The ductility of
plain carbon steels can drop precipitously at the transition temperature.
Figure 4.15 shows the Charpy impact strength at low temperatures for several
common structural materials. Note that in general the fcc and hcp materials retain
their resistance to impact, whereas the bcc materials tend to become brittle. Curve
6 of Fig. 4.15 shows the effect of a solid–solid phase transition around 140 K.

Figure 4.12 Tensile elongation and impact energy of a chromium–nickel stainless steel (AISI
316, annealed).

© 2005 by Marcel Dekker


5367-4 Flynn Ch04 R2 090904

268 Chapter 4

Figure 4.13 Tensile elongation and impact energy of a chromium stainless steel (AISI 430,
annealed).

5. LOW-TEMPERATURE STRENGTH OF SOLIDS

In general, the ultimate tensile strength of a solid material is greater at low tempera-
tures than it is at ordinary temperatures. This is true for both crystalline and
noncrystalline solids and for many heterogeneous materials as well (e.g., glass-
reinforced plastics). For metals, the strength at 4 K may be 2–5 times that at room
temperature. For plastics, the strength at 76 K may be 1.5–8 times greater than
the room-temperature value. Glasses show less change in strength at low tempera-
tures; at 76 K, glasses have between 1.5 and 2 times their room-temperature strength.
Reduction of the thermal energy of the metal lattice at lower temperatures is
responsible for part of the increase in the strength of metals. The decreased thermal
vibration at low temperatures also results in stronger plastics, because it becomes
more difficult for the long-chain molecules to slide over one another.
The change in strength of glass at low temperatures is not as great as that of
metals and plastics. The important cohesive forces of glasses are the chemical bonds
between atoms (rather than the intermolecular forces of plastics). Such chemical
bond forces are less affected by temperature.
There are exceptions to the above general rule. The thermal and mechanical
history of a material, various solid-state phase changes, the form of stress system,
and the rate of application of this stress system can also affect mechanical properties.

Figure 4.14 Tensile elongation and impact energy of plain carbon steels.

© 2005 by Marcel Dekker


5367-4 Flynn Ch04 R2 090904

Mechanical Properties of Solids 269

Figure 4.15 The Charpy impact strength at low temperatures for several common structural
materials. (1) 2024 T4 aluminum; (2) beryllium copper; (3) K Monel; (4) titanium; (5) 304
stainless steel; (6) C1020 carbon steel; (7) 9% nickel steel.

Some of the exceptions to the general rule of increasing strength with decreasing
temperature can be explained qualitatively by phase changes during deformation.
Some austenitic stainless steels undergo a partial transformation under strain from
a face-centered cubic to an intermediate hexagonal close-packed phase and then to
a body-centered cubic lattice (martensite). The martensite form is stronger than
the parent structure and contributes to the high strength of these materials. The ten-
sile strength of austenitic steels shows the usual increase with decreasing temperature
to some low temperature and then a decrease, reflecting a maximum in the phase
transformation curve.

5.1. Metals
The yield strength of ductile metals increases with decreasing temperature. The yield
strength at 20 K is between one and three times the room-temperature yield strength
for fcc and hcp metals. An even greater tendency toward increases in yield strength
at low temperatures is shown for bcc metals. In fact, for body-centered cubic steels,
the yield strength increases so rapidly with decreasing temperature that it becomes
greater than the fracture strength. Thus, below a certain temperature, these materials
fracture before they reach their yield strengths.
The yield and ultimate tensile strengths of annealed copper are shown in
Fig. 4.16. This figure shows that the yield strength of a material can be insensitive
to temperature while its tensile strength is increasing by more than a factor of 2.
The material has ‘‘work hardened.’’ That is, the material has hardened itself by
generating obstacles to dislocation motion.
The behavior of a plain carbon (bcc) steel is shown in Fig. 4.17 for contrast.
This behavior is typical of the metals that become brittle at low temperatures. The
yield and tensile strength curves approach one another as temperature is reduced.
When the two curves converge, the material is brittle in the tensile test.

© 2005 by Marcel Dekker


5367-4 Flynn Ch04 R2 090904

270 Chapter 4

Figure 4.16 Tensile and yield strength of annealed oxygen-free copper.

5.2. Plastics
Only a few plastics have been tested at temperatures below 200 K. Of these, only
Teflon showed ductility down to the lowest test temperature, which was 4 K. How-
ever, reinforced plastics such as the glass fiber laminates can have good properties,
the tensile strength parallel to the laminations increasing at low temperatures and
the modulus being approximately constant. Mylar breaks with fragmentation in a
tensile test and so is obviously brittle. Yet in films about 2.540  105 m (0.001 in.)
or less in thickness, Mylar shows remarkable flexibility in bending tests at tempera-
tures as low as 20 K.
Figure 4.18 shows the tensile strength of several plastics. The increase in
strength as the temperature is depressed is accompanied by a rapid decrease in elon-
gation and impact resistance. The glass-reinforced plastics are the only plastics that
retain appreciable impact resistance as the temperature is lowered. Figure 4.19 shows
the tensile strength and impact energy of a glass fiber-reinforced epoxy resin sample.
These glass-reinforced epoxies are very useful for cryogenic applications because of
their high strength-to-weight ratios and high strength-to-thermal conductivity ratios
(Weitzel et al., 1960).

5.3. Glass
The strength of glass at room temperature varies inversely with load duration, is
sensitive to atmospheric water vapor, and is also sensitive to rather minute surface

Figure 4.17 Tensile and yield strength of a plain carbon steel (SAE 1010 m, cold-rolled).

© 2005 by Marcel Dekker


5367-4 Flynn Ch04 R2 090904

Mechanical Properties of Solids 271

Figure 4.18 Tensile strength of selected plastics. (1) Polyethylene terephthalate (Mylar); (2)
polytetrafluoroethylene (Teflon); (3) polytrifluoromonochloroethylene (Kel-F); (4) polyvinyl
chloride; (5) nylon.

defects (Kropschot and Mikesell, 1957). The dependence of strength on load


duration (fatigue) has been found to decrease as the temperature is lowered below
ambient, but for a soda lime glass it was still appreciable at 83 K.
Table 4.2 shows the median breaking stress of borosilicate crown optical
glass (BSC-2) as a function of temperature. Owing to the brittle nature of glass,
there is large scatter in the breaking stress data; therefore, these data should not
be used for design purposes. For example, the median value at room temperature
(296 K) is 34.47 MPa (5000 psi) at a stress rate of increase of 1 psi=s. One percent
of the samples will fail at a stress of approximately 27.58 MPa (4000 psi) or
below. The strength properties of glass can be improved by tempering the
surface, i.e., placing the surface layer in compression. Large optical glass
(BSC-2) windows have been successfully tempered for use in liquid hydrogen
bubble chambers.

Figure 4.19 Tensile strength and impact energy of a glass fiber-reinforced epoxy resin.

© 2005 by Marcel Dekker


5367-4 Flynn Ch04 R2 090904

272 Chapter 4

Table 4.2 Breaking Stress of a Borosilicate Crown Optical Glass—Median Values from
Probability Plots

Breaking stress

Rate of stress increase 296 K 194 K 76 K 20 K

Condition psi=s MPa=s psi Mpa psi MPa psi MPa psi MPa

Abraded 800 5.52 7500 51.7 9500 65.5 10,400 71.71 10,400 71.71
Abraded 10 0.0689 5500 37.9 7500 51.71 10,400 71.17 10,600 73.08
Abraded 1 0.00689 5000 34.5 6400 44.13 10,400 71.17 10,200 70.33
Unabraded 800 10,400 71.7 18,000 124.11

5.4. Ultimate Stress


The ultimate stress (the maximum nominal stress attained during a simple tensile
test) is shown in Fig. 4.20 for several engineering materials. The yield stress of several
materials is shown in Fig. 4.21 (Durham et al., 1962). (The yield stress is the value of
stress at which the strain of the material in a simple tensile test begins to increase
rapidly with increase in stress or cause a permanent set of 0.1–0.2%.) At low
temperatures, less thermal agitation is available to assist dislocation motion, and
hence the yield stress usually increases.
Example 4.1. In many design situations, the yield strength-to-density ratio is an
important parameter. Determine the strength-to-density ratio Syr in Nm=kg for
the following materials at 22.2 K (density given in parenthesis):
a. 2024 T4 aluminum (2740 kg=m3)
b. 304 stainless steel (7506 kg=m3)
c. Monel (8885 kg=m3)

Figure 4.20 Ultimate stress for several engineering materials. (1) 2024 T4 aluminum; (2) ber-
yllium copper; (3) K Monel; (4) titanium; (5) 304 stainless steel; (6) C1020 carbon steel; (7) 9%
nickel steel; (8) Teflon.

© 2005 by Marcel Dekker


5367-4 Flynn Ch04 R2 090904

Mechanical Properties of Solids 273

Figure 4.21 Yield stress for the same engineering materials as in Fig. 4.20.

d. Beryllium copper (8304 kg=m3)


e. Teflon (2297 kg=m3)

From Fig. 4.21:


a. 2024 T4 aluminum

Sy ¼ 7:24  108 N=m2


Sy =r ¼ 2:64  N m=kg

b. 304 stainless steel

Sy ¼ 1:32  109 N=m2


Sy =r ¼ 1:76  105 N m=kg

c. Monel

Sy ¼ 1:03  109 N=m2


Sy =r ¼ 1:16  105 N m=kg

d. Beryllium copper

Sy ¼ 8:96  108 N=m2


Sy =r ¼ 1:08  105 N m=kg

e. Teflon

Sy ¼ 1:38  108 N=m2


Sy =r ¼ 6:00  104 N m=kg

© 2005 by Marcel Dekker


5367-4 Flynn Ch04 R2 090904

274 Chapter 4

6. ELASTIC CONSTANTS

Elastic constants are physical properties, which relate stress to strain, or force per
unit area to relative length change. Either stress or strain can be induced by any elas-
tically coupled force—mechanical, thermal, magnetic, or electrical. Since solids resist
both volume change and shape change, they have at least two independent elastic
constants. Most materials considered in this chapter are quasi-isotropic; their macro-
scopic elastic constants do not depend on direction. They are characterized elasti-
cally by two independent elastic constants. Composites and textured aggregates
are anisotropic rather than quasi-isotropic. They are characterized by five indepen-
dent elastic constants for the transverse-isotropic case and nine constants for the
orthotropic case. This section defines and describes the elastic constants used most
often to characterize polycrystals.

6.1. Compressibility, KT
All matter—gas, liquid, or solid—responds to pressure change and exhibits at least
one elastic constant, the compressibility. It is given by
KT ¼ ð1=V Þð@V =@PÞT ð4:1Þ
where P, T, and V denote pressure, temperature, and volume. Figure 4.22 illustrates
the mechanical deformations with which compressibility is associated. KT is always
positive and has units of reciprocal pressure.
kT ¼ ðA=‘Þðd‘=dLÞT ð4:2Þ
where A denotes the cross-sectional area of a rod-shaped, hydrostatically loaded spe-
cimen, l denotes length, and L denotes load. For isotropic solids, linear and bulk
compressibilities relate according to
3k ¼ KT ð4:3Þ

6.2. Bulk Modulus, B


The bulk modulus, B, is used to describe a solid’s resistance to volume change rather
than compressibility. For isotropic materials, B and K relate reciprocally:
BT ¼ 1=KT ¼ V ð@P=@V ÞT ð4:4Þ

Figure 4.22 Mechanical deformations with which compressibility or bulk modulus is


associated.

© 2005 by Marcel Dekker


5367-4 Flynn Ch04 R2 090904

Mechanical Properties of Solids 275

B has pressure units, and, herein, elastic constants with pressure units are called
elastic moduli. For describing solids, hydrostatic stress, s, is often preferred over
pressure, which is a negative stress. Thus,
BT ¼ V ð@s=@V ÞT ð4:5Þ
A high bulk modulus usually implies strong interatomic forces, high cohesive energy,
high melting point, and high elastic-stiffness moduli.

6.3. Young’s Modulus, E


Engineers find Young’s modulus the most familiar elastic constant, as evidenced by
its many pseudonyms: extension modulus, tensile modulus, tension modulus, elastic
modulus, and modulus. Young’s moduli in tension or compression are identical,
theoretically and experimentally.
Usually defined in terms of a uniaxially stressed rod where both stress, s, and
strain, e, are measured along the rod axis, z. Young’s modulus, E, becomes (see
Fig. 4.1)
E ¼ sz =ez ð4:6Þ
Figure 4.23 illustrates the mechanical deformations with which Young’s modulus is
associated.
As shown in Fig. 4.23, rod bending also involves Young’s modulus. The stress
along a bent rod is
sz ¼ E; ez ¼ Ex=r ð4:7Þ
where x is a coordinate perpendicular to z, and r is the curvature radius of the
neutral surface near the origin.

6.4. Shear Modulus, G


Sometimes called the torsional modulus, rigidity modulus, or transverse modulus,
the shear modulus, G, relates the shear stress, t in simple shear to the shear
strain, g. Thus,
t ¼ Gg ð4:8Þ

Figure 4.23 Mechanical deformations with which Young’s modulus is associated.

© 2005 by Marcel Dekker


5367-4 Flynn Ch04 R2 090904

276 Chapter 4

Figure 4.24 Mechanical deformations with which shear modulus and torsional modulus are
associated.

Figure 4.24 illustrates the mechanical deformations with which the shear
modulus is associated. A cube sheared on one face has a shear strain related to
the length change of the face diagonal by

g ¼ 2D‘=‘ ð4:9Þ

In torsion (see Fig. 4.24), t is the torsional stress. For isotropic materials, simple-
shear and torsional G’s are identical. Both shear and torsion conserve volume but
change shape. The bulk modulus, as described above, describes volume change
without shape change.

6.5. Poisson’s Ratio, n


Poisson’s ratio, n, is not an elastic modulus, but a dimensionless ratio of two elastic
compliances. Usually defined by a uniaxially stressed rod (Young’s-modulus) experi-
ment, Poisson’s ratio is the negative ratio of transverse (x) and longitudinal (z)
strains:
n ¼ ex =ez ð4:10Þ
Figure 4.25 illustrates the deformation and strains involved. A typical value for
metals is 1=3, 0.28–0.42 being the observed range for most materials. Such v values

Figure 4.25 Mechanical deformation and strains with which Poisson’s ratio is associated.

© 2005 by Marcel Dekker


5367-4 Flynn Ch04 R2 090904

Mechanical Properties of Solids 277

mean that a material tends to maintain constant volume during uniaxial deforma-
tion. Liquids exhibit constant volume during deformation, and some solids (such
as rubber) tend to.
The four elastic constants just described—B, E, G, and n—are sometimes
referred to as the bulk, engineering, and macroscopic, polycrystalline, practical, or
technological elastic constants. These four constants are those used most frequently
to describe materials such as polycrystalline aggregates.

6.6. Hooke’s Law


Hooke’s law states a linear proportionality between strain-response intensity, e, and
imposed stress, s
s ¼ Ce ð4:11Þ
where C denotes an elastic-stiffness modulus. For small strains, Eq. 4.11 is an
example of Hooke’s law written inversely:
e ¼ Ss ð4:12Þ
where the S’s are elastic-compliance moduli, and for isotropic media
C ¼ 1=S ð4:13Þ
This chapter considers mainly Hooke’s-law, or linear-elastic, solids. Nonlinear
cases involve elastic constants higher than second order and are not treated in this book.

6.7. Elastic-Constant Inter-relationships


Although for complete characterization isotropic solids require only two indepen-
dent elastic constants, practice requires many more. This multiplicity arises because
particular elastic constants best describe particular mechanical deformations. As
already discussed, B, E, and G best describe dilatation, extension, and shear, respec-
tively. Table 4.3 summarizes inter-relationships among various elastic constants.

6.8. Typical Elastic-Constant Values


Tables 4.4–4.7 give elastic constants at various temperatures for six metals—nickel,
iron, copper, titanium, aluminum, and magnesium—that form the bases for most

Table 4.3 Connecting Identities Among Isotropic-Solid Elastic Constants

B E G n
3BE 1 E
B, E – – 
9B  E 2 6B
9BG 1 3B  2G
B, G – – 
3B þ G 2 3B þ G
3B 1  2v
B, v – 3B (12v)  –
2 1þv
GE E
E, G – – 1
3ð3G  EÞ 2G
E E
E, v – –
3ð1  2vÞ 2ð1 þ vÞ
2G 1 þ v
G, v 2G (1 þ v) – –
3 1  2v

© 2005 by Marcel Dekker


5367-4 Flynn Ch04 R2 090904

278 Chapter 4

Table 4.4 Bulk Modulus (Reciprocal Compressibility) of Six Metals at Selected


Temperatures in Units of GPa

T (K) Ni Fe Cu Ti Al Mg

0 187.6 187.6 142.0 110.0 79.4 36.9


20 187.6 173.1 142.0 109.8 79.4 36.9
40 187.5 172.9 141.9 79.4 36.8
60 187.3 172.7 141.7 79.3 36.8
80 187.2 172.1 141.4 109.5 79.2 36.7
100 187.0 171.6 141.1 79.0 36.6
120 186.6 171.1 140.8 109.0 78.8 36.5
140 186.2 170.7 140.5 78.6 36.4
160 185.8 170.3 140.1 78.3 36.3
180 185.5 169.9 139.7 108.5 78.0 36.1
200 185.2 169.5 139.3 77.7 36.0
220 184.9 169.2 138.8 108.0 77.4 35.9
240 184.6 168.9 138.4 77.1 35.7
260 184.2 168.6 138.0 76.7 35.6
280 183.9 168.2 137.5 107.5 76.4 35.4
300 183.6 168.0 137.1 107.3 76.1 35.3

technological alloys. These data are taken from "Materials at Low Temperatures,"
Am. Soc. Metals (1983).
Table 4.8 gives ambient-temperature elastic constants for 24 cubic elements for
which reliable data exist. This table also includes mass density, r.
Table 4.9 gives zero-temperature values extrapolated from 4 K data. Since the
values shown in Tables 4.4–4.9 come from many different sources, some small differ-
ences in values will occur among the tables.
The data in Tables 4.4–4.9 are all taken from ‘‘Materials at Low Tempera-
tures,’’ Am. Soc. Metals, 1983.

Table 4.5 Young’s Modulus of Six Metals at Selected Temperatures in Units of GPa

T (K) Ni Fe Cu Ti Al Mg

0 240.1 224.1 138.6 130.6 78.4 49.4


20 240.1 224.1 138.6 130.6 78.3 49.3
40 239.8 223.9 138.4 78.2 49.2
60 239.2 223.5 138.0 77.9 49.1
80 238.5 222.9 137.5 129.4 77.6 48.9
100 237.5 222.3 136.9 77.1 48.6
120 236.4 221.7 136.2 127.0 76.5 48.3
140 235.1 220.9 135.4 75.8 48.0
160 233.8 220.1 134.6 75.2 47.6
180 232.7 219.3 133.7 123.7 74.5 47.2
200 231.3 218.4 132.9 73.8 46.8
220 230.0 217.7 132.0 120.0 73.1 46.4
240 228.6 216.8 131.1 72.3 46.0
260 227.2 215.8 130.1 71.6 45.6
280 225.8 214.9 129.2 116.6 70.8 45.1
300 224.5 214.0 128.2 114.6 70.1 44.7

© 2005 by Marcel Dekker


5367-4 Flynn Ch04 R2 090904

Mechanical Properties of Solids 279

Table 4.6 Shear Modules of Six Metals at Selected Temperatures in Units of GPa

T (K) Ni Fe Cu Ti Al Mg

0 93.3 87.3 51.9 50.2 29.3 19.4


20 93.3 87.2 51.8 50.2 29.3 19.3
40 93.2 87.2 51.8 29.3 19.3
60 93.0 87.0 51.6 29.2 19.1
80 92.6 86.8 51.4 49.7 29.0 19.1
100 92.2 86.6 51.2 28.8 19.0
120 91.8 86.4 50.9 48.8 28.6 18.9
140 91.2 86.1 50.5 28.3 18.8
160 90.6 85.7 50.2 28.1 18.6
180 90.1 85.4 49.9 47.2 27.8 18.4
200 89.6 85.0 49.6 27.5 18.3
220 89.0 84.7 49.2 45.7 27.2 18.1
240 88.4 84.3 48.8 26.9 17.9
260 87.8 83.9 48.5 26.6 17.7
280 87.2 83.5 48.1 44.2 26.3 17.6
300 86.6 83.1 47.7 43.4 26.0 17.4

7. MODULUS OF ELASTICITY

The constant of proportionality relating tensile stress and strain in the range of elas-
tic response of a material is the modulus of elasticity, or Young’s modulus, E. Two
other commonly used elastic moduli are (1) the shear modulus G, which is the rate of
change of shear stress with respect to shear strain at constant temperature in the
elastic region, and (2) the bulk modulus B, which is the rate of change of pressure
(corresponding to a uniform three-dimensional stress) with respect to volumetric
strain (change in volume per unit volume) at constant temperature.

Table 4.7 Poisson’s Ratio of Six metals at Selected Temperatures

T (K) Ni Fe Cu Ti Al Mg

0 0.287 0.285 0.338 0.302 0.336 0.277


20 0.287 0.285 0.338 0.302 0.336 0.278
40 0.287 0.285 0.338 0.336 0.278
60 0.287 0.285 0.338 0.336 0.278
80 0.288 0.285 0.338 0.303 0.337 0.279
100 0.288 0.284 0.338 0.337 0.279
120 0.289 0.284 0.339 0.306 0.338 0.280
140 0.290 0.285 0.339 0.339 0.281
160 0.290 0.285 0.340 0.340 0.282
180 0.291 0.285 0.341 0.311 0.341 0.283
200 0.292 0.285 0.341 0.342 0.284
220 0.293 0.286 0.342 0.315 0.343 0.285
240 0.294 0.286 0.342 0.344 0.286
260 0.294 0.287 0.343 0.345 0.287
280 0.295 0.287 0.343 0.321 0.346 0.288
300 0.296 0.288 0.344 0.322 0.347 0.289

© 2005 by Marcel Dekker


5367-4 Flynn Ch04 R2 090904

280 Chapter 4

Table 4.8 Room-Temperature Elastic Properties of 24 Cubic Elements

r (g=cm3) A B (GPa) E (GPa) G (Gpa) v

Ag 10.500 2.8825 101.20 78.89 28.79 0.3701


Al 2.697 1.2231 75.86 70.27 26.11 0.3456
Au 19.300 2.8522 173.50 76.12 26.68 0.4269
C 3.512 1.2109 442.00 1139.37 532.23 0.0704
Cr 7.200 0.7144 161.87 278.38 114.72 0.2134
Cu 8.930 3.2727 135.30 121.69 45.07 0.3501
Fe 7.872 2.4074 166.93 205.87 79.52 0.2945
Ge 5.323 1.6644 75.02 130.30 53.82 0.2105
Ir 22.520 1.5645 370.37 556.74 222.79 0.2495
K 0.851 6.7143 3.33 2.25 0.81 0.3874
Li 0.532 8.5243 12.13 9.21 3.35 0.3734
Mo 10.228 0.7140 259.77 321.83 124.40 0.2935
Na 0.971 7.1624 6.51 4.81 1.74 0.3788
Nb 8.570 0.4925 163.78 103.07 36.94 0.3951
Ni 8.910 2.6652 185.97 212.27 81.03 0.3098
Pb 11.344 4.0743 44.76 23.17 8.19 0.4136
Pd 12.038 2.8096 193.06 128.23 46.15 0.3893
Pt 21.500 1.5938 282.70 176.23 63.12 0.3961
Rb 1.560 7.4545 2.60 1.85 0.67 0.3811
Si 2.331 1.5636 97.89 161.61 65.97 0.2248
Ta 16.626 1.5609 189.72 183.69 68.61 0.3386
Th 11.694 3.6212 57.70 69.83 26.89 0.2983
V 6.022 0.7867 155.57 129.11 47.41 0.3617
W 19.257 1.0114 310.38 409.83 160.10 0.2799

If the material is isotropic (many polycrystalline materials can be consi-


dered isotropic for engineering purposes), these three moduli are related through
Poisson’s ratio m, the ratio of strain in one direction due to a stress applied perpen-
dicular to that direction to the strain parallel to the applied stress (see also
Table 4.3):

E
B¼ ð4:14Þ
3ð1  2mÞ

(in some texts, the symbol K is used for the bulk modulus),

E
G¼ ð4:15Þ
2ð1 þ mÞ

2Gð1 þ mÞ
E¼ ð4:16Þ
1þm

where

E  2G

2G

© 2005 by Marcel Dekker


5367-4 Flynn Ch04 R2 090904

Mechanical Properties of Solids 281

Table 4.9 Zero-Temperature Elastic Properties of 24 Cubic Elements

vt
r B E G vm vl (cm ms1)
(g=cm3) A (GPa) (GPa) (GPa) v (cm ms1) (cm ms1) y (K)

Ag 10.634 2.9912 108.72 87.02 31.84 0.3666 0.1950 0.3770 0.1730 226.46
Al 2.732 1.2073 79.38 78.24 29.29 0.3357 0.3674 0.6583 0.3274 430.56
Au 19.478 2.8436 180.32 83.19 29.23 0.4231 0.1391 0.3355 0.1225 161.74
C 3.516 1.2109 442.00 1139.37 532.23 0.0704 1.3416 1.8099 1.2304 2239.62
Cr 7.231 0.6865 201.00 300.58 120.16 0.2508 0.4526 0.7068 0.4076 589.57
Cu 9.024 3.1904 142.03 132.51 49.28 0.3445 0.2625 0.4798 0.2337 344.40
Fe 7.922 2.3431 169.00 213.17 82.64 0.2898 0.3603 0.5936 0.3230 472.44
Ge 5.340 1.6651 76.52 132.92 54.90 0.2105 0.3544 0.5295 0.3207 373.36
Ir 22.659 1.5698 366.67 557.12 223.43 0.2468 0.3485 0.5416 0.3140 429.62
K 0.910 7.6267 3.66 3.10 1.14 0.3587 0.1261 0.2386 0.1120 90.52
Li 0.5375 9.2400 13.30 10.70 3.90 0.3650 0.3049 0.5868 0.2706 326.74
Mo 10.223 0.9024 265.29 335.72 130.22 0.2891 0.3481 0.6552 0.3569 474.53
Na 1.005 8.5000 7.24 5.98 2.20 0.3623 0.1664 0.3181 0.1478 147.43
Nb 8.616 0.5183 173.03 110.87 39.79 0.3932 0.2430 0.5122 0.2149 276.60
Ni 8.972 2.3859 187.60 233.52 90.33 0.2925 0.3541 0.5860 0.3173 475.98
Pb 11.593 3.8379 48.79 30.34 10.86 0.3964 0.1095 0.2336 0.0968 105.33
Pd 12.092 2.4552 195.43 134.72 48.63 0.3851 0.2266 0.4640 0.2005 275.92
Pt 21.578 1.4828 288.40 183.54 65.83 0.3939 0.1976 0.4175 0.1747 238.44
Rb 1.629 8.0657 3.21 2.39 0.87 0.3761 0.0824 0.1638 0.0730 55.31
Si 2.331 1.5640 99.22 163.21 66.57 0.2258 0.5916 0.8980 0.5344 648.87
Ta 16.754 1.6154 194.21 191.15 71.54 0.3360 0.2319 0.4158 0.2066 263.77
Th 11.888 3.4545 58.10 75.58 29.45 0.2832 0.1754 0.2862 0.1574 163.83
V 6.050 0.8130 157.04 135.23 49.85 0.3565 0.3230 0.6078 0.2870 399.19
W 19.313 0.9959 314.15 417.76 163.40 0.2784 0.3240 0.5248 0.2909 384.39

or
3K  E

6K
It has been found that Poisson’s ratio for isotropic materials does not change
appreciably with change in temperature in the cryogenic range. Therefore, all three
of the elastic moduli above vary in the same manner with temperature. Accordingly,
in the discussion that follows, ‘‘modulus’’ shall mean Young’s modulus E, which is
the slope of the initial portion of the stress–strain curve of Fig. 4.1.
We expect materials to get stiffer at low temperatures, and that is the usual
case. However, the change in modulus of elasticity with temperature is not very great
in metals compared to the change in strength over the same temperature range. Some
general statements can be made regarding the temperature dependence of the elastic
modulus.
The elastic moduli of most polycrystalline metals increase by about 10% in
going from room temperature to about 20 K, and below this there is little change
(Campbell, 1974). For some alloys, the modulus of elasticity may either decrease
or increase in value or even exhibit maxima or minima as the temperature is lowered,
depending upon their composition. The moduli of noncrystalline plastics are 2–20
times their room temperature values at 4 K, whereas the moduli of glasses, also

© 2005 by Marcel Dekker


5367-4 Flynn Ch04 R2 090904

282 Chapter 4

Figure 4.26 Young’s modulus of fine common plastics.

noncrystalline, either increase or decrease by about 3% between room temperature


and 76 K, depending upon composition. Thermodynamic arguments predict that
the elastic constants of all solids will be independent of temperature at absolute zero,
so no great changes are to be expected in the moduli below about 20 K.
Nonmetallic materials exhibit a much greater change in the modulus of elasti-
city at low temperatures. For polyethylene terephthalate, E increases by a factor of
nearly 2 as the temperature is decreased from 300 to 76 K, and the modulus of poly-
tetrafluorethylene at 4 K is about 20 times that at room temperature (Fig. 4.26)
(Corruccini, 1957).
Curves of the modulus of elasticity of representative engineering materials are
shown in Fig. 4.27 (Durham et al., 1962). Similar information is summarized in
Table 4.10 (McClintock and Gibbons, 1960).

8. FATIGUE STRENGTH

A simple reverse bending test is the usual method used to measure fatigue strength.
Fatigue strength is defined as the stress required to cause failure after a certain num-
ber of bending cycles and is given as sf.

© 2005 by Marcel Dekker


5367-4 Flynn Ch04 R2 090904

Mechanical Properties of Solids 283

Figure 4.27 Young’s modulus of several metals at low temperatures. (1) 2024 T4 aluminum;
(2) beryllium copper; (3) K Monel; (4) titanium; (5) 304 stainless steel; (6) C1020 carbon steel;
(7) 9% nickel steel; (8) Teflon.

Table 4.10 Temperature Dependence of Modulus of Elasticity of Some Common Metals

Metals exhibiting very small change in Metals exhibiting very large change in
modulus of elasticity from 300 to 20 K modulus of elasticity from 300 to 20 K

Aluminum Aluminum
356 1100
7075 2024
Cobalt
Elgiloy
Copper
Berylco 25
Pure coppera
Iron Iron
Vascojet 1000a 17–7 pH
2800 (9% Ni) 304, 310
4340a 321, 347
Nickel Nickel
Inconel K Monel
Inconel Xa Pure nickel
Titanium Titanium
A-110-ATa C-210-AV
B-120-VCAa
a
No significant temperature dependence.

© 2005 by Marcel Dekker


5367-4 Flynn Ch04 R2 090904

284 Chapter 4

Figure 4.28 Fatigue strength of several materials at 106 cycles. (1) 2024 T4 aluminum; (2)
beryllium copper; (3) K Monel; (4) titanium; (5) 304 stainless steel; (6) C1020 carbon steel;
(7) 9% nickel steel; (8) Teflon.

Some materials, such as carbon steel, have the property that fatigue failure will
not occur if the stress is maintained below a certain value, called the endurance limit
se no matter how many cycles have elapsed.
Fatigue strength data at low temperatures are not as common as data on yield
strength and ultimate tensile strength, because the fatigue tests are more time-
consuming and hence more expensive to perform. As one would expect, however,
the data that have been reported all show that fatigue strength increases as tempera-
ture decreases, in the same manner as the yield and ultimate tensile strength.
Fortunately, for aluminum alloys at least, it has been found that the ratio of
fatigue strength to ultimate strength remains fairly constant as the temperature is
lowered. Therefore, the fatigue strength of these alloys varies with temperature in
the same manner that the ultimate strength varies with temperature. This fact may
be used in estimating the fatigue strength for nonferrous materials at cryogenic
temperatures if no fatigue data are available.
Representative data of fatigue strength of several materials at low temperature
are shown in Fig. 4.28.

9. MECHANICAL PROPERTIES SUMMARY

It is convenient to classify metals by their lattice structure for low-temperature


mechanical properties. The face-centered cubic (fcc) metals and their alloys are most
often used in the construction of cryogenic equipment. Aluminum, copper, nickel,
their alloys, and the austenitic stainless steels of the 18–8 type are fcc and do not
exhibit an impact ductile-to-brittle transition at low temperatures. As a general rule,
the structural properties of these metals improve as the temperature is reduced. The
yield strength at 22.2 K (40 R) is quite a bit greater than at ambient temperature;

© 2005 by Marcel Dekker


5367-4 Flynn Ch04 R2 090904

Mechanical Properties of Solids 285

Young’s modulus is 5–20% greater at the lower temperatures, and fatigue properties
are also improved at the lower temperatures (with the exception of 2024 T4 alumi-
num). Annealing of these metals and alloys can affect both the ultimate and yield
strengths.
The body-centered cubic (bcc) metals and alloys are usually undesirable for
low-temperature construction. This class of metals includes iron, the martensitic
steels (low-carbon and the 400 series of stainless steels), molybdenum, and niobium.
If not already brittle at room temperature, these materials exhibit a ductile-to-brittle
transition at low temperatures. Working of some steels, in particular, can induce the
austenite-to-martensite transition.
The hexagonal close-packed (hcp) metals exhibit structural properties inter-
mediate between those of the fcc and bcc metals. For example, zinc suffers a duc-
tile-to-brittle transition, whereas zirconium and pure titanium do not. Titanium
and its alloys, having an hcp structure, remain reasonably ductile at low tempera-
tures and have been used for many applications where weight reduction and reduced
heat leakage through the material have been important. Small impurities of O, N, H,
and C can have a detrimental effect on the low-temperature ductility properties of Ti
and its alloys.
Plastics also increase in strength as the temperature is decreased, but this is
accompanied by a rapid decrease in elongation in a tensile test and a decrease in
impact resistance as the temperature is lowered. The glass-reinforced plastics also
have high strength-to-weight and strength-to-normal conductivity ratios. Conver-
sely, all elastomers become brittle at low temperatures. Nevertheless, many of these
materials, including rubber, Mylar, and nylon, can be used for static seal gaskets
provided they are highly compressed at room temperature prior to cooling.
The strength of glass under constant loading also increases with decreases in
temperature. Since failure occurs at a lower stress when the glass has surface defects,
the strength can be improved by tempering the surface.
See Figs. 4.29–4.32 for the temperature dependence of several important
module: E.

10. DESIGN CONSIDERATIONS

A cryogenic storage container must be designed to withstand forces resulting from


the internal pressure, the weight of the contents, and bending stresses. Material com-
patibility with low temperatures, which has already been discussed, results in choos-
ing among the fcc metals (copper, nickel, aluminum, stainless steels, etc.). Because
these materials are more expensive than ordinary carbon steels, a design goal is to
make the inner vessel as thin as possible. This constraint also reduces cooldown time
and the amount of cryogenic liquid required for cooldown. Thus, the inner vessel of
cryogenic containers is nearly always thin walled.
A thin-walled cylinder has a wall thickness such that the assumption of
constant stress across the wall results in negligible error. Cylinders having
internal diameter-to-thickness (D=t) ratios greater than 10 are usually considered
thin-walled. An important stress in thin-walled cylinders is the circumferential, or
hoop, stress (see Fig. 4.33), which is

s ¼ pr=t ð4:17Þ
where, p is the internal pressure, r the radius and, t the thickness of cylinder.

© 2005 by Marcel Dekker


5367-4 Flynn Ch04 R2 090904

286 Chapter 4

Figure 4.29 Temperature variation of (E) Young’s modulus for six metals.

Figure 4.30 Temperature variation of bulk modulus (B)(reciprocal compressibility) for six
metals.

© 2005 by Marcel Dekker


5367-4 Flynn Ch04 R2 090904

Mechanical Properties of Solids 287

Figure 4.31 Temperature variation of shear modulus (G) for six metals.

Figure 4.32 Temperature variation of Poisson’s ratio (v) for six metals.

© 2005 by Marcel Dekker


5367-4 Flynn Ch04 R2 090904

288 Chapter 4

Figure 4.33 Hoop stress in a thin-walled cylinder.

If the cylinder is closed at the ends, a longitudinal stress of pr=2t is developed.


These basic equations of mechanics have been transformed into design codes
such as the ASME Boiler and Pressure Vessel Code, Section VIII (1983) and the
British Standards Institution Standard 1500 or 1515. According to the ASME Code,
Section VIII, the minimum thickness of the inner shell for a cylindrical vessel should
be determined from
PD PDo
t¼ ¼ ð4:18Þ
2sa ew  1:2P wsa ew þ 0:8P

where P is the design internal pressure (absolute pressure for vacuum-jacketed


vessels), D the inside diameter of shell, Do the outside diameter of shell, sa the
allowable stress (approximately one-fourth minimum ultimate strength of material)
and ew is the weld efficiency.
Referring to Fig. 4.33, it can be seen that the tensile stress developed in a thin
hollow sphere is

s ¼ Pr=2t
The ASME Code that transforms this equation for practical use requires that
the minimum thickness for spherical shells, hemispherical heads, elliptical heads, or
ASME torispherical heads be determined from the equation
PDK PDo K
th ¼ ¼ ð4:19Þ
2sa ew  0:2P 2sa ew þ 2PðK  0:1Þ

where D is the inside diameter of the spherical vessel or hemispherical head, the
inside major diameter for an elliptical head, or 2 (crown radius) for the ASME
torispherical head. The value of the constant K is given by

1

6½2 þ ðD=D1 Þ2 

where D1 is the minor diameter of the elliptical head. For the ASME torispherical
head, K ¼ 0.885.
The outer shell, which has a high vacuum on one side and atmospheric pressure
on the other, is subject to the failure mode of external pressure collapse. A
thin-walled external shell will collapse under an externally applied pressure at stress
values much lower than the yield strength for the material.

© 2005 by Marcel Dekker


5367-4 Flynn Ch04 R2 090904

Mechanical Properties of Solids 289

To determine the collapsing pressure, the outer shell is assumed to be perfectly


round and of uniform thickness, the material obeys Hook’s law, the radial stress is
negligible, and the normal stress distribution is linear. Other assumptions are also
made. Using the theory of elasticity, the collapsing pressure is
 t 3
Pc ¼ KE ð4:20Þ
D
The factor K, a numerical coefficient, depends upon the L=D and D=t ratios (D is the
outer shell diameter), the kind of end support, and whether pressure is applied
radially only or at the ends as well. This failure mode is covered by the ASME Code,
in which design charts are presented for the design of cylinders and spheres
subjected to external pressure.
One simple embodiment of the collapsing pressure for a long cylinder exposed
to external pressure is given by

2Eðt=Do Þ3
Pc ¼ ð4:21Þ
1  m2
where E is the Young’s modulus of shell material, t the shell thickness Do the outside
diameter of shell and m the Poisson’s ratio for shell material.
A long cylinder in this analysis has an L=D such that

L=Do > 1:140ð1  m2 Þ1=4 ðDo =tÞ1=2 ð4:22Þ


where L is the unsupported length of the cylinder (distance between stiffening rings
for the outer shell). When the cylinder is stiffened with rings, the shell may be
assumed to be divided into a series of shorter shells equal in length to the ring
spacing.
For short cylinders, the collapsing pressure may be estimated from

2:42Eðt=Do Þ5=2
Pc ¼ ð4:23Þ
ð1  m2 Þ3=4 ½L=Do  0:45ðt=Do Þ1=2 
where L is the distance between stiffening rings for the outer shell.
The heads that enclose the ends of the outer shell cylinder are also subject to
collapsing pressure. The critical pressure for a hemispherical, elliptical, or torisphe-
rical head (or for a spherical vessel) is given by

0:5Eðth =Ro Þ2
Pc ¼ ð4:24Þ
½3ð1  m2 Þ1=2
where th is the thickness of the head and Ro is the outside radius of the spherical
head, or the equivalent radius for the elliptical head, or the crown radius of the
torispherical head. The equivalent radius for elliptical heads is given by Ro ¼ K1D,
where D is the major diameter and K1 is the shape factor.
Stresses on the piping needed to fill and empty cryogenic containers must also
be taken into account.
The minimum wall thickness for piping under internal pressure is given by the
ASA Code as
PDo

2sa þ 0:8P

© 2005 by Marcel Dekker


5367-4 Flynn Ch04 R2 090904

290 Chapter 4

where P is the design pressure, Do the outside diameter of pipe, and sa the allowable
stress of pipe material.
For piping subjected to external pressure, Eq. (4.21) may be used.
In addition to pressure-induced stress, thermal stresses on the piping are often
a serious consideration. When the deformation arising from change of temperature is
prevented, temperature stresses arise that are proportional to the amount of defor-
mation that is prevented. Let a be the coefficient of expansion per degree of tempera-
ture, l1 the length of bar at temperature T1, and l2 the length at temperature T2. Then

l2 ¼ l1 ½1 þ aðT2  T1 Þ

If, subsequently, the structured member is cooled to a temperature T1, the propor-
tionate deformation is s ¼ a(T2  T1), and the corresponding unit stress is

s ¼ E a xðT2  T1 Þ:

This discussion of design considerations is by no means complete. Load-


induced bending stresses, flexure stresses, and support systems are not mentioned
at all. The intent here is merely to show the practical application of the low-tempera-
ture properties discussed in this section. For complete design considerations, one
should consult Mark’s Standard Handbook for Mechanical Engineers (1978), the
appropriate codes, and such classic strength of materials books as that of
Timoshenko and Gere (1961).
Example 4.2. A food freezing firm is currently using an ammonia refrigeration sys-
tem to maintain a food warehouse temperature of 255.6 K. Plans are being drawn to
change to a spray liquid nitrogen system in which controlled amounts of liquid nitro-
gen are sprayed directly into the cold storage room. Because of this, it must be
assumed that at times parts of the warehouse shelving will drop to 77.8 K.
You must determine whether or not the existing structures can be safely used.
Assume that the shelving was assembled at 300 K and that

Coefficient of linear expansion b ¼ 1.656  107= C
Modulus of elasticity E ¼ 1.86  1013 Pa
Yield strength ¼ 2.76  108 Pa
Ultimate tensile strength ¼ 1.17  109 Pa
Working stress ¼ 1=5  ultimate strength
a. Will the shelves rupture if exposed to liquid nitrogen?
b. Will permanent set occur?
c. From the working stress viewpoint, is the design satisfactory?
d. For the next generation plant, a quick choice must be made between alu-
minum and brass shelving. Is there any obvious low-temperature material
reason (not considering cost) to prefer one over the other at this stage of
the design?
By definition,
s stress
E¼ ¼ ¼ 1:86  1013 Pa ð4:25Þ
E strain
Dl Dl E
b¼ ; E¼ ¼b¼ or E ¼ bDT
lDT l DT

© 2005 by Marcel Dekker


5367-4 Flynn Ch04 R2 090904

Mechanical Properties of Solids 291

Stress:
s ¼ EE ¼ bDTE
¼ ð1:656  107 Þð255:6  77:8Þð1:86  1013 Þ ð4:26Þ
¼ 5:48  108 Pa

a. Stress < ultimate strength of 1.17  109 Pa. Therefore, material will not
rupture.
b. Stress > yield strength of 2.76  108 Pa. Therefore, permanent set will
occur. Not satisfactory.
c. Working stress ¼ (1=5)(1.17  109 Pa) ¼ 2.34  108 Pa.
Stress 5.48 108 Pa > 2.34  108 Pa.
Stress exceeds working stress. Not satisfactory.
d. No. Aluminum and brass are both fcc metals and should serve equally
well, all other things being the same.

11. MATERIAL SELECTION CRITERIA FOR CRYOGENIC TANKS


11.1. General Design Considerations
High-performance cryogenic storage vessels employ the concept of the dewar flask
design principle—a double-walled container with the space between the two vessels
evacuated and filled with an insulation. A fill line and a drain line are essential
elements for larger vessels, and these may be one and the same. A vapor vent line
must also be provided to allow the vapor formed as a result of heat inleak to escape
and prevent overpressurization of the vessel.
Cryogenic fluid storage vessels are not designed to be filled completely for two
major reasons. First, heat leak into the container is always present; therefore, the
vessel pressure would rise quite rapidly if little or no vapor space were allowed.
Second, inadequate cooldown of the inner vessel during a rapid filling operation
would result in excessive boil-off, and the liquid would be percolated through the
vent tube and wasted if no ullage space were provided. A 10% ullage volume is
commonly used for large storage vessels.
The details of conventional cryogenic fluid storage vessel design are covered in
such standards as the ASME Boiler and Pressure Vessel Code, Section VIII (1983)
for unfired vessels. Most users require that the vessels be designed, fabricated, and
tested according to this code.
The inner vessel must be constructed of a material compatible with the cryo-
genic fluid. Therefore 18-8 stainless steel, aluminum, Monel, Inconel, and titanium
are commonly used for the inner shell. These materials are much more expensive
than ordinary carbon steel, so the designer would like to make the inner vessel wall
as thin as possible to hold the cost within reason. In addition, a thick-walled vessel
requires a longer cooldown time, wastes more liquid in cooldown, and introduces the
possibility of thermal stresses in the vessel wall during cooldown.
Table 4.11 lists the allowable stresses for some of these materials.
As discussed earlier, an important property of metals chosen for cryogenic ser-
vice is their grain size. Grain boundaries, which are discontinuities in the orientation
of neighboring crystals, are obstacles to slip-plane movement. A metal having a
small grain size will have a higher yield stress than one with a larger grain size.

© 2005 by Marcel Dekker


5367-4 Flynn Ch04 R2 090904

292 Chapter 4

Table 4.11 Allowable Stress for Materials at Room Temperature or Lower

Allowable Stress

Material MPa psi

Carbon steel (for outer shell only) SA-285 Grade C 94.8 13,750
SA-299 129.2 18,750
SA-442 Grade 55 94.8 13,750
SA-516 Grade 60 103.4 15,000
Low-alloy steel SA-202 Grade B 146.5 21,250
SA-353-B (9% Ni) 163.7 23,750
SA-203 Grade E 120.6 17,500
SA-410 103.4 15,000
Stainless steel SA-240 (304) 129.2 18,750
SA-240 (304L) 120.6 17,500
SA-240 (316) 129.2 18,750
SA-240 (410) 112.0 16,250
Aluminum SB-209 (1100-0) 16.2 2,350
SB-209 (3004-0) 37.9 5,500
SB-209 (5083-0) 68.9 10,000
SB-209 (6061-T4) 41.4 6,000
Copper SB-11 46.2 6,700
SB-169 (annealed) 86.2 12,500
Nickel alloys (annealed) SB-127 (Monel) 128.2 18,600
SB-168 137.9 20,000

Source: ASME Code, Sec. VIII (1983).

One reason annealed metals are more ductile than work-hardened ones is that work
hardening increases the grain size.
For large storage vessels, 9% nickel steels are commonly used with higher boil-
ing cryogens (T > 75 K), whereas aluminum alloys and austenitic steels can generally
be used over the entire range of liquid temperatures. To speed up cooldown time and
lower the cost of the cooldown of the container, the thickness of the inner shell is
kept as small as possible and is only designed to withstand the maximum internal
pressure and the bending forces. Section VIII of the ASME Boiler and Pressure
Vessel Code provides all the necessary design equations and should be adhered to
for safety reasons when designing vessels of capacities greater than 0.2 m3.

11.2. Constitution and Structure of Steel


As a result of the methods of production, the following elements are always present in
steel: carbon, manganese, phosphorus, sulfur, silicon, and traces of oxygen, nitrogen,
and aluminum. Various alloying elements are frequently added, such as nickel, chro-
mium, copper, molybdenum, and vanadium. The most important constituent is carbon,
and it is necessary to understand the effect of carbon on the internal structure of steel.
When pure iron is heated to 910 C, its internal crystalline structure changes from a
body-centered cubic arrangement of atoms (alpha-iron) to face-centered cubic (gamma-
iron). The alpha-iron containing carbon, or any other element in solid solution, is called
ferrite; and the gamma-iron containing other elements in solid solution is called austenite.
When the iron solution is cooled rapidly, and the carbon does not have time to
separate out in the form of carbide, the austenite transforms to a highly stressed

© 2005 by Marcel Dekker


5367-4 Flynn Ch04 R2 090904

Mechanical Properties of Solids 293

structure supersaturated with carbon. This structure is called martensite. Martensite


is exceedingly hard, but it is brittle and requires tempering to increase the ductility.
Therefore, martensite should be avoided for cryogenic service.

11.3. Nickel Steel and Nickel Alloys


Nickel steel is steel that contains nickel as the predominant alloying element. Some
nickel steels are the toughest structural materials known. Compared to steel, other
nickel steels have ultrahigh strength, high proportional limits, and high moduli of
elasticity. At cryogenic temperatures, nickel alloys are strong and ductile.
Nickel added to carbon steel increases the steel’s ultimate strength, elastic limit,
hardness, and toughness. It narrows the hardening range but lowers the critical range
of steel, reducing the danger of warping and cracking, and balances the intensive
deep-hardening effect of chromium. Nickel steels are also of finer structure than
ordinary steels, and the nickel retards grain growth.
When the percentage of nickel is high, the steel is very resistant to corrosion.
Above 20% nickel, the steel becomes a single-phase austenitic structure. The steel
is nonmagnetic above 29% nickel, and the maximum value of magnetic permeability
is at about 78% nickel. The lowest coefficient of thermal expansion value of the steel
is at 36% nickel.
The fracture and toughness properties of steels have been extensively studied.
The results consistently indicate that the 5–9% nickel steels retain satisfactory
properties down to 260 F. As the temperature is reduced from 260 F to
320 F, the fatigue and fracture resistance may be degraded for the 5% nickel
steel. Therefore, for temperatures below 320 F, such as liquid hydrogen tempera-
tures, a 9% nickel steel is a minimum requirement. The data in Table 4.12 give some

Table 4.12 Relative Values for the Properties of Nickel Alloys

Thermal
Young’s Shear Thermal expansion Specific
Temperature Density modulus modulus Poisson’s conductivity (mean) heat
( F) (lb=in3) (104 psi) (106 psi) ratio (Btu=ft=h= F) (106= F) (101Btu=lb= F)

3.5Ni
70 0.284 29.6 11.5 0.282 20 6.6 108
150 – 30.5 11.9 0.281 17 5.7 84
5Ni
70 0.282 28.7 11.2 0.283 18 6.6 108
260 – 30.1 11.8 0.277 12 5.2 60
320 – 30.3 11.8 0.277 9.2 4.9 36
9Ni
70 0.283 28.3 10.7 0.286 16 6.6 108
260 – 29.6 11.2 0.281 10 5.2 60
320 – 29.8 11.3 0.280 7.5 4.9 36
36Ni (Invar)
70 0.282 22.1 8.08 0.284 8.0 0.65
320 – 20.4 7.37 0.307 3.6 1.00
452 – 20.6 7.33 0.332 0.18 0.77

© 2005 by Marcel Dekker


5367-4 Flynn Ch04 R2 090904

294 Chapter 4

relative values for the properties of nickel alloys at both room and cryogenic
temperatures.
Commercial nickel and nickel alloys are available in a wide range of wrought
and cast grades; however, considerably fewer casting grades are available. Wrought
alloys tend to be better known by trade names such as Monel, Hastelloy, Inconel,
and Incoloy. There is also a class of superalloys that are nickel-based, strengthened
by intermetallic compound precipitation in a face-centered cubic matrix.
Most wrought nickel alloys can be hot- and cold-worked, machined, and
welded successfully. The casting alloys can be machined or ground, and many can
be welded or brazed. Nearly any shape that can be forged in steel can be forged
in nickel and nickel alloys. However, because nickel work-hardens easily, severe
cold-forming operations require frequent intermediate annealing to restore soft
temper. Annealed cold-rolled sheet is best for spinning and other manual work.

11.3.1. Behavior in Welding


Nickel alloys can be joined by shielded metal arc, gas tungsten arc, gas metal arc,
plasma arc, electron beam, oxyacetylene, and resistance welding; by silver and
bronze brazing; and by soft soldering. Resistance welding methods include spot,
seam, protection, and flash welding.
Special nickel alloys, including superalloys, are best worked at about 800–
2200 F. In the annealed condition, these alloys can be cold-worked by all standard
methods. Required forces and rate of work hardening are intermediate between
those of mild steel and Type 304 stainless steel. These alloys work-harden to a
greater extent than the austenitic stainless steels, so they require more intermediate
annealing steps.

11.4. Stainless Steels


The principal reason for the existence of stainless steels is their outstanding resis-
tance to corrosion. However, stainless steels also exhibit excellent mechanical prop-
erties, and in some applications these mechanical properties are as important as
corrosion resistance in determining the life or utility of a given structure.
Within the stainless steel family, there are martensitic, ferritic, austenitic, and
precipitation-hardenable steels. These steels exhibit a wide variety of mechanical
properties at room and cryogenic temperatures. The martenstic and precipitation-
hardenable steels can be heat-treated to high strengths at room temperature. Thus,
these steels are frequently used for structural applications. The ferritic stainless steels
exhibit better resistance to corrosion than the martensitic stainless steels and can be
polished to a high luster. However, they cannot be treated to high strengths. Thus,
they are used in applications where appearance is more important, such as in auto-
motive and appliance trim.

11.5. Austenitic Stainless Steels


The austenitic stainless steels exhibit a wide range of mechanical properties. They are
soft and ductile in the annealed condition and can be easily formed. They can also be
appreciably strengthened by cold-working, but even at high strength levels they exhi-
bit good ductility and toughness. This good ductility and toughness are retained to
cryogenic temperature. Thus, austenitic steels are used in the annealed condition in

© 2005 by Marcel Dekker


5367-4 Flynn Ch04 R2 090904

Mechanical Properties of Solids 295

applications requiring good formability. In the cold-worked condition, they are


frequently used in structural applications. For cryogenic temperature applications,
the austenitic steels are unsurpassed because they are relatively easy to fabricate
and weld, do not require heat treatment after fabrication, and exhibit relatively
high strength with excellent stability and toughness at very low temperatures.
Austenitic chromium–nickel stainless steels are the most widely used stainless
steels. While resistance to corrosion is their chief attribute, they also display low
magnetic permeability, good high-temperature strength, and excellent toughness at
low temperatures.
The variations in composition among the standard austenitic stainless steels are
important, both in the performance of the steel in service and in its behavior in welding.
As examples, Types 302, 304, and 304L represent the so-called 18-8 stainless steels. They
differ chiefly in carbon content, which determines the amount of carbide precipitation
that can occur in the base metal heat-affected zones near a weld. Types 316 and 316L
contain an addition of molybdenum for improved corrosion resistance. The presence
of molybdenum will have both advantages and disadvantages in welding.
The austenitic steels are very well suited to cryogenic service. Consider the effect
of cryogenic temperatures on the tensile properties of the austenitic stainless steels
shown in Table 4.13 (ASM, 1983). This table lists the tensile properties of four auste-
nitic stainless steels used in cryogenic service at room temperature, 320 F, and
425 F: Types 304, 304L, 310, and 347. Note that the high ductility (elongation and
reduction of area) of the austenitic stainless steels is retained at cryogenic temperatures.
Note also that the yield and tensile strengths of these steels increase as the temperature
decreases, with the largest increase occurring in the tensile strength. The one exception
is Type 310, which exhibits almost the same increase in tensile strength as in yield
strength. This behavior is probably due to the slightly higher amount of carbon in
Type 310 compared to the other steels listed. Interstitial elements, such as carbon,
are known to have marked effects on the yield strength exhibited at low temperatures.
As indicated by the ductility exhibited by the austenitic steels, the toughness of
these steels is excellent at cryogenic temperatures. The results of Charpy V-notch
impact tests on Types 304, 304L, 310, and 347 conducted at room temperature,

Table 4.13 Tensile Properties of Some Austenitic Stainless Steels

Testing Yield strength Tensile strength Elongation in Reduction in


AISI type temp. ( F) (0.2 % offset) (psi) (psi) 2 in. (%) area (%)

304 75 33,000 85,000 60 70


304 –320 57,100 205,500 43 45
304 –425 63,700 244,500 48 43
304L 75 28,000 85,000 60 60
304L 320 35,000 194,500 42 50
304L 425 33,900 220,000 41 57
310 75 45,000 95,000 60 65
310 –320 84,900 157,500 54 54
310 –425 115,500 177,500 56 61
347 75 35,000 90,000 50 60
347 –320 41,200 186,000 40 32
347 –425 45,500 210,500 41 50

Source: ASM (1983).

© 2005 by Marcel Dekker


5367-4 Flynn Ch04 R2 090904

296 Chapter 4

Table 4.14 Transverse Charpy V-Notch Impact


Strength of Some Austenitic Stainless Steels

Energy absorbed (ft lb)


AISI
type 80 F 320 F 425 F

304 154 87 90
304L 118 67 67
310 142 89 86
347 120 66 57

Source: ASM (1983).

320 F, and 425 F are shown in Table 4.14. The toughness of these four steels
decreases somewhat as the temperature is decreased from room temperature to
320 F, but on the further decrease to 425 F the toughness remains about the
same. It should be noted that the toughness of these steels is excellent at 425 F
and that the toughness of Types 304 and 310 is significantly higher than those of
Types 304L and 347. Shown in Table 4.15 are the results of impact tests on the same
four steels at 320 and 425 F for various thickness of plate. These results demon-
strate that the toughness of the austenitic stainless steels is not markedly affected by
plate thickness.
Table 4.16 shows the results of impact tests done on Type 304 stainless steel
after up to 1 year of exposure at 320 F. These results indicate that Type 304 is very

Table 4.15 Low-Temperature Impact Strength of Several Annealed Austenitic Stainless


Steels

Testing temp Specimen Type of Energy


AISI type ( F) orientation notch Product size absorbed (ft lb)

304 –320 Longitudinal Keyhole 3 in. plate 80


304 –320 Transverse Keyhole 3 in. plate 80
304 –320 Transverse Keyhole 21=2 in. plate 70
1
304 –425 Longitudinal Keyhole =2 in. plate 80
304 –425 Longitudinal V-notch 31=2 in. plate 91.5
304 –425 Transverse V-notch 31=2 in. plate 85
1
304L 320 Longitudinal Keyhole =2 in. plate 73
1
304L 320 Transverse Keyhole =2 in. plate 43
304L 320 Longitudinal V-notch 31=2 in. plate 67
304L 425 Longitudinal V-notch 31=2 in. plate 66
310 –320 Longitudinal V-notch 31=2 in. plate 90
310 –320 Transverse V-notch 31=2 in. plate 87
310 –425 Longitudinal V-notch 31=2 in. plate 86.5
310 –425 Transverse V-notch 31=2 in. plate 85
1
347 –320 Longitudinal Keyhole =2 in. plate 60
1
347 –320 Transverse Keyhole =2 in. plate 47
347 –425 Longitudinal V-notch 31=2 in. plate 59
347 –425 Transverse V-notch 31=2 in. plate 53
347 –300 Longitudinal V-notch 61=2 in. plate 77
347 –300 Transverse V-notch 61=2 in. plate 58

© 2005 by Marcel Dekker


5367-4 Flynn Ch04 R2 090904

Mechanical Properties of Solids 297

Table 4.16 Charpy Keyhole Impact Values at –320 F


on Type 304 After Prolonged Exposure at 320 F

Time of exposure Energy absorbed (ft lb)

0 85.0
30 min 72.5
6 months 73.0
12 months 75.0

stable and does not exhibit any marked degradation of toughness due to long-time
exposure at 320 F.

11.5.1. Behavior in Welding


The austenitic stainless steels are considered the most weldable of the high alloy
steels. This favorable appraisal is mainly based on the great toughness of welded
joints, even in the as-welded condition. Comparatively little trouble is experienced
in making satisfactory welded joints if the inherent physical characteristics and
mechanical properties of an austenitic steel are given proper consideration. For
example, the austenitic stainless steels have a coefficient of expansion approximately
50% greater than that of plain steel, while the thermal conductivity is about
one-third as great. This calls for more attention to factors that control warpage
and distortion and to thermally induced stresses.
An important part of the successful welding of austenitic stainless steels is the
control of compositions and microstructures through selection of type, welding pro-
cedure, and postweld treatment. As the steels become more complex in composition
or heavier in section or the service conditions become more demanding, a greater
degree of knowledge of stainless steel metallurgy is needed.
For instance, in the heat-affected zone adjacent to the welds in austenitic stain-
less steels, chromium carbides will precipitate to varying degrees. This precipitation
of carbides, known as sensitization, depends on the carbon content of the stainless
steel and the exposure time in the range of 800–1600 F. Because these carbides pre-
cipitate in the grain boundaries, they may be detrimental to the toughness of the
steels. However, the toughness of the lower carbon Type 304 and 304L steels remains
at relatively high levels after sensitization. Thus, for cryogenic equipment that is
welded, low-carbon austenitic stainless steels such as Type 304 appear to be better
materials of construction than the high-carbon austenitic stainless steels such as
Type 302.

11.6. Titanium Alloys


Depending on the predominant phase or phases in their microstructure, titanium
alloys are categorized as alpha, beta, and alpha–beta. The alpha phase in pure tita-
nium is characterized by a hexagonal close-packed crystalline structure that remains
stable from room temperature to approximately 1620 F. The beta phase in pure tita-
nium has a body-centered cubic structure and is stable from 1620 F to the melting
point, 3040 F.

© 2005 by Marcel Dekker


5367-4 Flynn Ch04 R2 090904

298 Chapter 4

The addition of alloying elements to titanium provides a wide range of physical


and mechanical properties. Some elements, notably tin and zirconium, behave as
neutral solutes and have little effect except to strengthen the alpha phase.
Like stainless steel, titanium sheet and plate work-harden significantly during
forming. Minimum bend-radius rules are nearly the same for both, although spring-
back is greater for titanium. Commercially pure grades of heavy plate are cold-
formed or, for more severe shapes, warm-formed at temperatures of about 800 F.
Alloy grades can be formed at temperatures as high as 1400 F in inert gas
atmospheres.
Titanium plates can be sheared, punched, or perforated on standard equip-
ment. The harder alloys are more difficult to shear, so thickness limitations are
generally about two-thirds of those for stainless steel. Titanium and its alloys can
be machined and abrasive-ground; however, sharp tools and continuous feed
are required to prevent work hardening. Tapping is difficult because the metal galls.
Coarse threads are used whenever possible.

11.6.1. Welding Titanium Alloys


Generally, titanium is welded by gas tungsten arc (GTA) or plasma arc techniques.
Inert gas processes can be used under certain conditions. In all aspects, GTA welding
of titanium is similar to that of stainless steel. Normally, a sound weld appears bright
silver with no discoloration on the surface or along the heat-affected zone.

11.7. Aluminum and Aluminum Alloys


Nonferrous metals such as aluminum offer a wide variety of mechanical properties
and material characteristics. In addition to property variations, nonferrous metals
and alloys differ in cost, based on availability, abundance, and the ease with which
the metal can be converted into useful forms.
When selecting any material for a mechanical or structural application, some
important considerations include how easily the material can be shaped into a fin-
ished product and how its properties change or are altered, either intentionally or
inadvertently, in the shaping process. Like steels, nonferrous metals can be cast,
rolled, formed, forged, extruded, or worked by other deforming processes. Although
the same operations are used with ferrous and nonferrous metals and alloys, the
reaction of nonferrous metals to these forming processes is often more severe.
Consequently, properties may differ considerably between cast and wrought forms
of the same metal or alloy. Each forming method imparts unique physical and
mechanical characteristics to the final component.
Aluminum and its alloys, numbering in the hundreds, are available in all com-
mon commercial forms. Aluminum alloy sheet can be formed, drawn, stamped, or
spun. Many wrought or cast aluminum alloys can be welded, brazed, or soldered,
and aluminum surfaces readily accept a wide variety of finishes. Aluminum
reflects radiant energy throughout the entire spectrum, is nonsparking, and is
nonmagnetic.
Though light in weight, commercially pure aluminum has a tensile strength of
about 13,000 psi. Cold working the metal approximately doubles it strength. In other
attempts to increase strength, aluminum is alloyed with elements such as manganese,
silicon, copper, magnesium, or zinc. The alloys can also be strengthened by cold
working. Some alloys are further strengthened and hardened by heat treatments.

© 2005 by Marcel Dekker


5367-4 Flynn Ch04 R2 090904

Mechanical Properties of Solids 299

At subzero temperatures, aluminum is stronger than at room temperature and is no


less ductile. However, most aluminum alloys lose their strength at elevated
temperatures.
Wrought aluminum alloys are the most useful for low-temperature service. A
four-digit number that corresponds to a specific alloying element usually designates
wrought aluminum alloys (Guyer, 1989).
To develop strength, heat-treatable wrought aluminum alloys are solution
heat-treated, then quenched and precipitation-hardened. Solution heat treating
consists of heating the metal, then holding it at temperature to bring the hardening
constituents into solution, then cooling to retain those constituents in solution.
Precipitation hardening after solution heat treatment increases the strength and
hardness of these alloys.
Wrought aluminum alloys are also strengthened by cold working. The high-
strength alloys, whether heat-treatable or not, work-harden more rapidly than the
softer, lower strength alloys and so may require annealing after cold working.
Because hot forming does not always work-harden aluminum alloys, this method
is used to avoid annealing and straightening operations. However, hot forming fully
heat-treated materials is difficult. Generally, aluminum formability increases with
temperature.

© 2005 by Marcel Dekker


5367-4 Flynn Ch05 R2 090904

5
Transport Properties of Solids

1. THERMAL PROPERTIES

The thermal properties of materials at low temperatures of most interest to the


process engineer are specific heat, thermal conductivity, and thermal expansivity.
(We shall use the term ‘‘specific heat’’ to refer to a material property and the term
‘‘heat capacity’’ to refer to the atomic scale contributions to specific heat.) It will
be shown that each of these properties depends on the intermolecular potential of
the lattice and accordingly that they are interrelated.

1.1. Specific Heat


1.1.1. Lattice Heat Capacity
Nearly all the physical properties of a solid (e.g., specific heat, thermal expansion)
depend on the vibration or motion of the atoms in the solid. Specific heat is often
measured at low temperatures for design purposes. However, specific heat measure-
ments are important in their own right, because the variation of specific heat with
temperature shows how energy is distributed among the various energy-absorbing
modes of the solid. Thus specific heat measurements give important clues to the
structure of the solid. Finally, because other properties also depend on the lattice
structure and its vibration, specific heat measurements are used to predict or corre-
late other properties, such as thermal expansion. Therefore, an understanding of the
temperature dependence of specific heat not only gives useful design information but
also is helpful in predicting other thermal properties.
The specific heat of any material is defined from thermodynamics as
 
@U
CV ¼ ð5:1Þ
@T V
where U is the internal energy, T is the absolute temperature, and V is the volume.
CV is the property more useful to theory than CP, because it directly relates
internal energy, and hence the microscopic structure of the solid, to temperature.
However, it must be remembered that most solids expand when they are heated at
constant pressure. As a result, the solid does work against both internal and external
forces. The specific heat measured at constant pressure, CP, then, includes some
extra heat to provide this work. Under ordinary circumstances, CP is the specific heat
observed. Therefore, CV must be calculated from CP. Fortunately, using thermody-
namic theory, this is not difficult, and the relation between CP and CV will be shown
301

© 2005 by Marcel Dekker


5367-4 Flynn Ch05 R2 090904

302 Chapter 5

later. In addition, for solids (and liquids), the difference between CP and CV is
usually less than 5% at room temperature. Accordingly, we shall limit the present
discussion to CV, to lattice effects only, and to electrical insulators. The specific heat
contribution of free electrons in a metal will be taken up after the lattice effects have
been explored.
(a) Models of Lattice Specific Heat. The various models of lattice specific heat
will be discussed in the order of their historical development, which is also more or
less the order of their exactness.
Dulong and Petit. In 1911, Dulong and Petit observed that the heat capacity
of many solids, both metals and nonmetals, was very nearly independent of tempera-
ture near room temperature, and furthermore that it had a value of 6 cal=(mol K )
regardless of the substance. This experimentally observed fact can be explained if
it is assumed that each lattice point can absorb energy in the same way as every other
lattice point (the equipartition of energy). In this model, the atoms are assumed to
oscillate as independent classical particles. Thus the total thermal energy of 1 mol
of material would be

 
kT
U ¼ ð3Þð2ÞNo ¼ 3No kT ð5:2Þ
2

where 3 is the number of degrees of freedom, 2 arises from the fact that the bound
oscillator has equal amounts of both kinetic and potential energy, No is Avogadro’s
number, the number of particles per mole, and (kT=2) is the thermal energy of each
degree of freedom.
From Eq. (5.1), then,

@U
CV ¼ ¼ 3No k ¼ 6 cal=ðmol KÞ
@T

which is the Dulong–Petit value.


This brief derivation indicates that CV is not a function of temperature, but this
is not the case. Instead, heat capacity varies with temperature as shown in Fig. 5.1
and goes to zero as the absolute temperature goes to zero.
Einstein. The theory of lattice specific heat was basically solved by Einstein,
who introduced the idea of quantized oscillation of the atoms. He pointed out that,
owing to the quantization of energy, the law of equipartition must break down at
low temperatures. Improvements have since been made on this model, but they all
include the quantization. Einstein treated the solid as a system of simple harmonic
oscillators of the same frequency. He assumed that each oscillator is independent.
This is not really the case, but the results, even with this assumption, were remark-
ably good. All the atoms are assumed to vibrate, due to their thermal motions, with a
frequency n, and according to quantum theory each of the three degrees of freedom
has an associated energy E¼hn=(ehn=kT1) in place of the kT postulated by classical
mechanics. The lattice specific heat, CV1, thus becomes temperature-dependent.
Einstein deduced for the specific heat at constant volume the formula
 
dU eyE =T ðyE =TÞ2
CV 1 ¼ ¼ 3R ð5:3Þ
dT V ðeyE =T  1Þ2

© 2005 by Marcel Dekker


5367-4 Flynn Ch05 R2 090904

Transport Properties of Solids 303

Figure 5.1 Specific heat curve typical of simple isotropic solids.

or
 2 !
yE eyE =T
CV 1 ¼ 3R ð5:4Þ
T ðeyE =T  1Þ2

where Nok has been replaced by its equivalent R, the general gas constant, and where
yE ¼ hn=k, which has the dimensions of temperature and is the Einstein characteristic
temperature. The characteristic temperature yE is the parameter that permits varia-
tion from one material to another. It permits plotting all materials on one curve by
replacing T by T=yE as the abscissa. Expression (5.4) fits experimental data quite well
for all materials except that at low temperatures it drops below the experimental
data. At intermediate and high temperatures, the fit is good, and Eq. (5.4) provides
the approach to a limiting Dulong–Petit value of 3R at high T. Although now super-
seded by more exact models, the Einstein model included the most important effect,
the quantization.
Nernst and Lindemann. The Einstein model was a considerable improvement
over that of Dulong and Petit, but it still did not have the correct low-temperature
shape. Real atoms in the lattice do not vibrate at only one frequency, but are coupled
to one another in a complex way that depends upon both the lattice type and the
nature of the interatomic forces.
Nernst and Lindemann improved upon the Einstein model by assuming that
the atoms vibrate at two frequencies, one equal to the Einstein frequency n, and
one at half that value, n=2. This treatment, while an improvement, was still a
considerable oversimplification and was on shaky theoretical grounds because it
required the introduction of a half quantum of energy.
Debye. Debye made a major advance in the theory of heat capacity at low
temperatures by treating a solid as an infinite elastic continuum and considering
the excitement of all possible standing waves in the material. These range from
the acoustic vibrations up to a limiting frequency, n m. Instead of a system of simple

© 2005 by Marcel Dekker


5367-4 Flynn Ch05 R2 090904

304 Chapter 5

harmonic oscillators all vibrating at only one frequency (Einstein) or two frequencies
(Nernst and Lindemann), Debye derived a parabolic frequency distribution for the
atoms vibrating in the lattice. The Debye model gives the following expression for
the lattice heat capacity per mole (CV1):
 3 Z yD =T 4 x  3  
T x e dx T T
CV 1 ¼ 9R 2
¼ 3R D ð5:5Þ
yD 0 ðex  1Þ yD yD

where R is the universal gas constant per mole, T is the absolute temperature, yD is
the Debye characteristic temperature, and x is a dimensionless variable defined by
the expression x ¼ hn=kT. In this last expression, h is Planck’s constant, n is the
frequency of vibration, and k is Boltzmann’s constant.
In the abbreviated form of the equation, D(T=yD) stands for the expression in
the integral, the Debye function. The Debye expression for CV vs. T=yD is plotted in
Fig. 5.1.
The important thing to notice about the Debye function is that for a given
substance, the lattice heat capacity is dependent only on a mathematical function
of the ratio of the Debye characteristic temperature yD to the absolute temperature.
This mathematical function is the same for all materials, and only yD changes from
material to material. yD thus appears as a constant characteristic of the material at
hand.
The value of yD is an adjustable parameter for the temperature scale, similar to
yE for the Einstein model. Its value usually lies between 200 and 400 K, although
values occur far outside this range. Values of yD are given in Table 5.1.
The Debye function predicts heat capacity surprisingly well, especially consid-
ering that Debye merely assumed a parabolic form for the vibration spectrum of
every solid without regard for the individual differences to be expected among
different solids.
At high temperatures (T > 2yD), the specific heat given by Eq. (5.5) approaches
a constant value of 3R, which is the Dulong–Petit value, as it should be. At low tem-
peratures (T < yD=12), the Debye function approaches a constant value of
D(0) ¼ 4p4=5; thus, the lattice specific heat at temperatures less than yD=12 may be
expressed as

CV 1 ¼ ð12p4 R=5ÞðT=yD Þ3 ¼ 464:5ðT=yD Þ3 ¼ AT 3 ð5:6Þ

or

CV ¼ AT 3 ð5:7Þ

where A is a constant.
From Eq. (5.6), we see that the lattice contribution to the specific heat of solids
varies as the third power of the absolute temperature at very low temperatures. This
behavior is useful in separating the effects of electrons from those of the lattice for
electrical conductors, and this will be used later.
The Debye temperature, yD, is not only important to estimating heat capacity
but is also a useful property in itself. yD can be calculated directly from first
principles by the theoretical expression
 
hn a 3N 1=3
yD ¼ ð5:8Þ
k 4pV

© 2005 by Marcel Dekker


Table 5.1 Debye Characteristic Temperatures (yD) of Some Representative Elements and Compounds at T  yD=2

Transport Properties of Solids


yD yD yD yD
   
Element R K Element R K Element R K Compound R K

Ac 180 100 Ge 666 370 Pb 153 85 AgCl 324 180


Ag 396 220 H (para) 207 115 Pd 495 275 Alums 144 80
Al 693 385 H (ortho) 189 105 Pr 216 120 As2O3 252 140
Ar 162 90 H (n-D2) 189 105 Pt 405 225 As2O5 432 240
As 495 275 He 54 30 Rb 108 60 AuCu3 (ord.) 360 200
Au 324 180 Hf 351 195 Re 540 300 AuCu3 (disord.) 324 180
B 2196 1220 Hg 180 100 Rh 630 350 BN 1080 600

5367-4 Flynn Ch05 R2 090904


Be 1692 940 I 189 105 Rn 720 400 CaF2 846 470
Bi 216 120 In 252 140 Sb 252 140 Cr2O3 648 360
C (diamond) 3650 2028 Ir 522 290 Se 270 150 FeS 1134 630
C (graphite) 1500 2700 K 180 100 Si 1134 630 KBr 324 180
Ca 414 230 Kr 108 60 Sn (fcc) 432 240 KCl 414 230
Cd (hcp) 504 280 La 234 130 Sn (tetra) 252 140 KI 351 195
Cd (bcc) 306 170 Li 756 420 Sr 306 170 LiF 1224 680
Ce 198 110 Mg 594 330 Ta 414 230 MgO 1440 800
Cl 297 165 Mn 756 420 Tb 315 175 MoS2 522 290
Co 792 440 Mo 675 375 Te 234 130 NaCT 504 280
Cr 774 430 N 126 70 Th 252 140 RbBr 234 130
Cs 81 45 Na 270 150 Ti 639 355 RbI 207 115
Cu 558 310 Nb 477 265 Tl 162 90 SiO2 (quartz) 459 255
Dy 279 155 Nd 270 150 V 504 280 TiO2 (rutile) 810 450
Er 297 165 Ne 108 60 W 567 315 ZnS 468 260
Fe 828 460 Ni 792 440 Y 414 230
Ga (rhomb) 432 240 O 162 90 Zn 450 250
Ga (tetra) 225 125 Os 450 250 Zr 432 240
Gd 288 160 Pa 270 150 AgBr 252 140

305
© 2005 by Marcel Dekker
5367-4 Flynn Ch05 R2 090904

306 Chapter 5

where h is the Planck’s constant; n a the speed of sound in the solid; k the
Boltzmann’s constant and N=V is the number of atoms per unit volume for the solid.
In practice, the Debye temperature is usually determined by selecting the value
that makes the theoretical specific heat curve fit the experimental specific heat curve
as closely as possible. In such cases, yD has lost its physical significance as a measure
of limiting frequency and has become an ‘‘effective’’ characteristic temperature.
A modification of the Debye theory is sometimes used to calculate heat
capacity in a semiempirical way. Essentially, yD is treated as a function of tempera-
ture rather than as a characteristic constant. A few heat capacities are measured, and
yD is allowed to vary with temperature to get the best fit between the Debye function
and experiment. Usually, the resulting variation of yD with temperature is not too
great for metals. Plastics, however, are characterized by large changes in their intera-
tomic force constants with temperature (equivalent to large changes in their elastic
moduli). As a result, yD for these materials increases greatly with decreasing
temperature.
The situation for compounds and alloys is more complicated. Near room
temperature, the specific heat of a compound or alloy can be approximated quite
well by linear combination of the specific heats of the constituent elements (the
Kopp-Joule rule). This is because the temperature is sufficiently high that all modes
of lattice vibration are fully excited (except for a few hard substances of low atomic
weight such as diamond, BeO, B2O3), and the gram atomic heat capacity is about the
same for all the constituent elements (the Dulong-Petit value).
This procedure is not proper at low temperatures because there the specific heat
is determined by the arrangement and bonding of the lattice, and for a compound
this bears no relation to the lattice characteristics of the component elements. This
is seen at once by comparing the solid properties of an alkali halide with those of
the alkali metal and the halogen as separate entities. Nevertheless the Debye function
is found to represent simple compounds that form lattices with only one type of
bonding (such as alkali halides, other binary salts, and some oxides). The yD value
in this case bears no relation to the yD value of the constituent elements.
Other Models. Although the model of Debye gives good agreement with
observation in many cases, deviations are observed. The cause seems to be that
the detailed crystal structure is not introduced. Much further work has been done,
beginning with Born and von Karman and Blackman and continued by many others.
These models attempt to express the detailed situation of actual crystal lattices. This
leads immediately to dispersion; that is, the velocity of sound depends upon the
frequency. In all these models, the phonon (quantized sound wave or lattice vibra-
tion) picture is retained, and it has become one of the well-established particles.
The frequency spectra differ from model to model, and each crystal leads to a differ-
ent frequency spectrum. These other models are very important to achieving a
detailed understanding of the crystal lattice. For process design purposes, however,
the Debye function and yD yield adequate estimations. As usual, though, good
experimental data are much to be preferred.
The specific heats of several materials used in low-temperature construction are
shown in Fig. 5.2 as a function of temperature. Values for other materials are tabu-
lated in several excellent references (Corruccini and Gniewek, 1960; Johnson, 1960;
Landolt-Bornstein, 1961), and a few values are abstracted in Table 5.2 (Gopal, 1966).

Example 5.1. Determine the lattice specific heat of chromium at 200 K as given by
the Debye function.

© 2005 by Marcel Dekker


5367-4 Flynn Ch05 R2 090904

Transport Properties of Solids 307

Figure 5.2 Specific heat of several materials of construction.

For chromium, yD ¼ 430 K from Table 5.1 and molecular weight ¼ 52


T=yD ¼ 200=430 ¼ 0.046
From Fig. 5.1, CV=R ¼ 2.5 and R ¼ 1.987 cal=(mol K)
CV ¼ ð2:5Þ½1:987cal=ðmol K=ð52g=molÞ
¼ 0:096 cal/(g K) (compares to published value of 0.099 cal/(g K), Table 5.2)
(b) Estimation Techniques. In the absence of experimental data, it may be
necessary to estimate heat capacity and hence yD. This may be done with modest
success for elements and simpler compounds.
Procedures for calculating values of yD, using either elastic constants, compres-
sibility, melting point, or the temperature dependence of the expansion coefficient,
and other properties, are outlined by Dillard and Timmerhaus (1968) and are
summarized briefly here.
Of all the properties that do depend on the lattice constants and hence reflect
yD, usually only the melting point and thermal expansion coefficient are likely to be
known better than the heat capacity itself. The formula relating melting point Tm
with yD is due to Lindemann:
 n 1=3 nT 1=2
m
yD ¼ 140 ð5:9Þ
V M
where V and M are molar volume (cm3) and molecular weight, respectively, and n is
the number of atoms in the molecule. Equation (5.9) has been shown to be a good
approximation for common metals and cubic binary salts.
An experimental study of alloys showed that the specific heat of the copper–
nickel alloys can be closely represented by using either the Kopp–Joule additive rule
or a weighted mean yD value, and it seems probable that generally this will be true
for simple alloy systems in which all compositions and the parent metals are of the
same lattice type and in which the yD values of the parent metals are not widely

© 2005 by Marcel Dekker


308
Table 5.2 Specific Heats (CP) of Some Selected Substances [cal=(g K)]

Temperature (K) Al Mg Cu Ni a-Mn a-Fe g-Fe Cr 18–l8 SSa Monel Fused silica Pyrex Teflon

5367-4 Flynn Ch05 R2 090904


20 0.0024 0.0040 0.0019 0.0012 0.0025 0.0011 0.0014 0.0006 0.0011 0.0014 0.006 0.0055 0.0183
50 0.0337 0.0580 0.0236 0.0164 0.0211 0.0129 0.0218 0.0090 0.016 0.0186 0.0272 0.0264 0.0491
77 0.0815 0.119 0.0471 0.0392 0.0473 0.0343 0.0487 0.0277 0.038 0.0417 0.0470 0.047 0.0739
90 0.102 0.141 0.0554 0.0488 0.0574 0.0441 0.0604 0.0381 0.050 0.0509 0.0570 0.0575 0.0851
100 0.116 0.155 0.0607 0.0555 0.0641 0.0516 0.0684 0.0459 0.057 0.0571 0.0643 0.065 0.0931
150 0.164 0.202 0.0774 0.0785 0.0872 0.0775 0.0975 0.0757 0.085 0.0782 0.0982 0.101 0.132
200 0.191 0.221 0.0854 0.0915 0.1003 0.0918 0.118 0.0925 0.099 0.0897 0.129 0.132 0.166
298 0.215 0.235 0.0924 0.1060 0.1146 0.1070 0.1251 0.1073 0.114 0.1019 0.177 0.182 0.248b
a
SS ¼ stainless steel.
b
At 280 K.

Chapter 5
© 2005 by Marcel Dekker
5367-4 Flynn Ch05 R2 090904

Transport Properties of Solids 309

different. The additive rule will fail for alloys having crystal structures and elastic
constants widely different from those of either of the major components.
It was noted that the Debye and other functions give CV, whereas one desires
CP for most purposes. The difference between CP and CV is given by several
thermodynamic expressions, one of which is
CP  CV ¼ a2 VT=bT ð5:10Þ
in which V is the molar volume, a is (1=V )(dV=dT )P, which is equal to three times
the coefficient of linear expansion, and bT is –(1=V )(dV=dP)T. However, these
coefficients are not always known. Since the difference between the two specific heats
is small, 1–10% of CP, it can often be neglected or calculated approximately from the
expression
CP  CV ¼ ACV 2 T ð5:11Þ
in which A ¼ a2V=bTCV 3 is regarded as a constant and is obtained from data at
normal temperatures.
If, to go one step further, the data are not available for calculating A at any
temperature from the above formula, A can be obtained to a less satisfactory degree
of accuracy from a formula by Nernst and Lindemann
A ¼ 0:0214=Tm ð5:12Þ
for specific heat in cal=(mol K) and Tm in K. Finally, it may be admitted that if CV
itself has to be estimated by an indirect method, the accuracy of the resulting heat
capacities will probably not be such as to justify taking CP – CV into account at all.
Often the property actually desired (as in heat balances) is the enthalpy differ-
ence over a large temperature interval, and this difference is fortunately much less
sensitive to estimation errors than is the specific heat.

1.1.2. Electron Heat Capacity


We have seen that the heat capacities of both metals and nonmetals can be well
predicted by considering only the energy of the lattice. The Dulong–Petit high-
temperature value is the same for both. This is paradoxical because the free electrons
in a metal should be capable of accepting some of the thermal energy of the metal,
and this demands that metals have a significantly larger heat capacity than nonme-
tallic crystals. This is, in fact, not the case, because only a small fraction of the
conduction electrons can accept thermal energy. The basic reason for this lies in
the Pauli exclusion principle, which limits the number of electrons that can occupy
the same energy state. All the possible low-lying energy states of electrons are already
occupied. Therefore, energy can be accepted only by those electrons that are in states
below the highest occupied energy level by an amount that corresponds to the
average energy of a thermal vibration. This energy is of the order of kT, where k
is Boltzmann’s constant. Therefore, only electrons within about kT of the highest
energy level occupied, the Fermi level, can be excited by thermal processes out of
the sea of already occupied states to a higher energy state. All others, since they have
less energy, cannot be excited to an unoccupied level by this amount of energy and so
are unable to accept it. This characteristic of the element is termed degeneracy.
Because the conduction electrons are degenerate until thermal energies characteristic
of temperatures on the order of 10,000 K are available, the contribution of conduc-
tion electrons to the specific heat of solids is small at ordinary temperatures.

© 2005 by Marcel Dekker


5367-4 Flynn Ch05 R2 090904

310 Chapter 5

Oddly enough, it is possible to observe the electron heat capacity of metals at


very low temperatures. Since theory predicts that only the electrons within an energy
range kT of the Fermi level of electron energy states will accept thermal energy from
the lattice, the electronic heat capacity will be linearly proportional to the absolute
temperature.
Meanwhile, at sufficiently low temperatures, Debye’s equation reduces to
 3
T
CV ¼ 464:5 ¼ constant  T 3
yD
See Eq. (5.6). This fits well the measured specific heat curves of many
substances, provided the temperature is below yD=12; for copper and aluminum
this is at about 30 K. On further reduction of the temperature to the boiling point
of helium, departures from the T3 law become evident and the specific heat is
given by an equation of the form
 
T 3
CV ¼ A þ gT ð5:13Þ
yD
The first term on the right is, as before, the contribution arising from the lattice
vibrations, and the second has been shown to be due to the specific heat of the
conduction electrons, which is negligible at high temperatures in comparison with
the lattice contribution and proportional to temperature at low temperatures. For
a typical metal, copper for example,

A ¼ 464:5; g ¼ 0:000177; yD ¼ 310 K

so that at 26 K, where T ¼ yD=12, the lattice specific heat is 0.275 cal=(g-atom) and
the electronic specific heat is only 0.0046 cal=(g-atom). At 4 K, the two contributions
to the specific heat are 0.001 and 0.007 cal=(g-atom), respectively, and at 1 K they are
0.0000155 and 0.000178 cal=(g-atom). Thus the electronic contribution, which is of
considerable theoretical interest, can most accurately be determined by measure-
ments at temperatures below about 4 K. This is a typical example of the contribu-
tions to physical theory that can be made as a result of low-temperature research.
The coefficient g in Eq. (5.13 ) is given by

4p4 ame MR2 T


CV e ¼ ¼ ge T ð5:14Þ
h2 No ð3p2 N=V Þ2=3

where a is the number of valence electrons per atom; me the electron mass; M the
atomic weight of material; R the specific gas constant for material; T the
absolute temperature; h the Planck’s constant; No the Avogadro’s number and
N=V is the number of free electrons per unit volume.
Example 5.2. Determine the ratio of the electronic contribution to the specific heat,
Cve, to the lattice specific heat (as determined by the Debye function) for copper at
(a) 5.6 K, (b) 0.56 K, and (c) 0.056 K.

4p4 ame MR2 T


yD ¼ 309:4 K; CV e ¼ ¼ ge T
h2 No ð3p2 N=V Þ2=3
ge ¼ 1:13  102 J=ðkg K2 Þ for copper; R ¼ 1:987 cal=ðmol KÞ

© 2005 by Marcel Dekker


5367-4 Flynn Ch05 R2 090904

Transport Properties of Solids 311

a. At 5.6 K:
J
CV e ¼ ð1:13  102 Þ ð5:6 KÞ ¼ 0:0645 J=ðkg KÞ
kg K2
T=yD ¼ 5:6=309:4 ¼ 0:0181
 3  
T 3 cal
CV ¼ 233:8 R ¼ 2:75  10 ð4184Þ ¼ 11:5 J=ðkg KÞ
yD gK
CV e =CV ¼ 5:61  103
b. At 0.56 K, by similar calculations,
CV e =CV ¼ 0:561
c. At 0.056 K,
CV e =CV ¼ 56:1

1.1.3. Anomalies in the Heat Capacity Curve


The Debye theory used to present the heat capacity–temperature relationship shown
in Fig. 5.1 depends on the way in which heat is accepted by the lattice and by
conduction electrons in metals. This model gives smooth functions of heat capacity
vs. temperature. Heat capacity measurements on many materials, however, do not
show such regular behavior. Sharp spikes are sometimes apparent in the heat capa-
city curves. Other mechanisms, besides those of conduction electrons and the lattice,
must come into play. Obviously, other mechanisms must exist to accept thermal
energy as the temperature is raised. These other energy-absorbing mechanisms
include phase transformations, Curie points, order–disorder transformations, rota-
tional transitions of molecules, transitions between spin states in paramagnetic salts,
and electronic excitations.
If these other mechanisms can accept energy, they will become active when the
thermal energy of the lattice approaches the activation energy of the new process.
That is, when kT approaches the order of the energy needed for the new process,
then the process can begin to accept additional heat. These phenomena can lead
to the peaks and bumps found in some materials.

1.2. Thermal Conductivity


The thermal conductivity Kt of any material is defined such that the heat transferred
per unit time dQ=dy is given by
dQ dT
¼ Kt A ð5:15Þ
dy dx
where A is the cross-sectional area and dT=dx is the thermal gradient.
To understand how thermal conductivity depends on temperature, especially at
low temperatures, it is useful to understand the basic mechanisms for energy
transport through materials.
There are three basic energy transports, and hence heat conduction, through a
solid:
1. By lattice vibrational energy transport, also called phonon conduction,
which occurs in all solids—dielectrics and metals. In nonmetallic crystals
and some intermetallic compounds, the principal mechanism of heat

© 2005 by Marcel Dekker


5367-4 Flynn Ch05 R2 090904

312 Chapter 5

conduction is by this lattice vibration mode or the mechanical interaction


between molecules. For single crystals at quite low temperatures, this
mode of heat conduction can be very effective, equaling or exceeding the
conduction by pure metals.
2. By electron motion, as in metals. Metals, of course, also have a lattice
structure and hence a lattice contribution to their thermal conductivity.
However, thermal conductivity in pure metals (particularly at low
temperatures) is due principally to the ‘‘free’’ conduction electrons, those
that are so loosely bound to the atoms that they wander readily throughout
the crystal lattice and thus transfer thermal energy.
3. By molecular motion, such as in organic solids and gases. This character-
istic disorder and the lattice imperfections of these materials introduce
resistance to heat flow. Accordingly, the disordered dielectrics such as glass
and polymeric plastics are the poorest solid conductors of heat.

For the present, we are concerned with structural materials, so this overview is
limited to dielectrics (insulators) and metals. Furthermore, since dielectrics have only
one heat transport mechanism (phonons) and metals have two (phonons plus
electrons), we shall begin by examining heat transport in dielectrics.

1.2.1. Dielectric Heat Conduction


We have already used the term phonon several times. In the discussion of thermal
conduction by the thermal vibrations of the crystal lattice, it has been found conve-
nient to treat the quantized vibrational modes as quasi-particles. Since these wave
packets have many of the properties of particles, they have been named phonons
in analogy with the photon, the ‘‘particle’’ of light. The phonon then is a ‘‘particle’’
of thermal energy. These quasi-particles undergo collisions with each other, just as
the molecules of a gas do, and the transfer of thermal energy in a perfect crystal
becomes closely analogous to the momentum transfer between gas molecules in
thermal agitation. A thermally excited solid is thus treated as a gas composed of
phonons.
Since the thermal energy of a solid can be treated in terms of phonons, the solid
becomes comparable to a box full of a phonon gas. In this model, the thermal
conductivity of a solid can be calculated in terms of the thermal conductivity of a
gas. For the latter, we can go to the classical kinetic theory of gases. This predicts
that heat flows across an imaginary surface, drawn perpendicular to a temperature
gradient, by hot (excess velocity) atoms passing through the surface plane from
one side to the other and by cold (low-velocity) atoms passing through from the
other side. This leads to the following expression for the thermal conductivity:
Kt ¼ ð1=3ÞCV UL ð5:16Þ
where CV is the specific heat of the phonons, the lattice specific heat; U the
velocity of propagation of the phonons, which travel at the speed of sound; and L
is the mean free path of the phonon between collisions.
The 1=3 is included because the phonons are in random motion and are free to
move in any of the three spatial directions.
This general expression for lattice thermal conductivity results directly from
the adoption of a kinetic model in which ‘‘particles’’ carry the heat. It can be
imagined that the capacity of a phonon to carry heat is CV; that, on the average,

© 2005 by Marcel Dekker


5367-4 Flynn Ch05 R2 090904

Transport Properties of Solids 313

one-third of the phonons are traveling in the desired direction, they travel at a
velocity U, and they travel a distance L before being diverted. To arrive at a knowl-
edge of how the thermal conductivity should vary with temperature, it is necessary to
consider the various mechanisms by which phonon scattering takes place and the
dependence upon temperature of each mechanism. This amounts to examining the
temperature dependence CV, U, and L (Eq. (5.16)).
The specific heat CV is the lattice specific heat of the solid, as already discussed.
Its variation with T is plotted in Fig. 5.3b. The average velocity U of the phonons is
the mean of the velocity of sound. U varies only slightly with temperature as plotted
in Fig. 5.3c. The phonon gas differs from a real gas in that the number of particles
varies with the temperature, increasing in number as temperature is increased.
At high temperatures, the large number of phonons leads to collisions between
phonons. Then, as T increases, L decreases, as shown in Fig. 5.3d.
At low temperatures, phonon–phonon collisions disappear and L is determined
by the distance between imperfections in the crystal. The phonons collide with
imperfections such as impurity atoms, dislocations, intercrystalline boundaries, or
finally, the specimen boundaries. The distance between these does not depend on
temperature, so L becomes constant at low T.
The general shape of the curve of K vs. T can now be readily expressed in terms
of the other curves. U is almost constant for all T, so it has negligible effect on the
shape. At low T, L is constant. Thus the K curve has the shape of the one remaining
variable, CV. At high T, CV is constant, so the K curve has the shape of the L curve,
which drops off with temperature. At intermediate temperature, a broad maximum
occurs in the transition between the high- and low-temperature behavior.
These considerations lead to the general temperature dependence of Kt shown
in Fig. 5.3a, which is experimentally observed also. For all materials, it starts at zero
for absolute zero temperature, rises to a maximum, and then falls asymptotically
toward zero as the temperature becomes very high.
Materials differ in the height of the maximum and its position in temperature,
as can be seen in Fig. 5.4, which shows the thermal conductivity curves for several
samples. The variability of thermal conductivity with impurity and cold work and
the suppression of the maximum in Kt is shown in Fig. 5.3. Note especially the ther-
mal conductivity of the dielectric, corundum. Its peak conductivity, around 37 K, is
almost 15 times that of copper at room temperature, making it an excellent thermal

Figure 5.3 Thermal conductivity mechanisms of dielectrics vs. temperature.

© 2005 by Marcel Dekker


5367-4 Flynn Ch05 R2 090904

314 Chapter 5

Figure 5.4 Thermal conductivity of several materials.

conductor. Glass is an extreme example of a disordered material and is shown for


comparison. In some cases, thermal conductivity becomes a function of specimen
dimensions.
It has been shown that thermal conductivity can be expressed in terms of the
Debye temperature yD and the thermal conductivity at the Debye temperature, K0:

Kt ¼ K0 ðyD =TÞ ð5:17Þ

This expression holds when T is equal to or greater than yD and gives a 1=T
dependence on temperature.
When T is considerably less than yD, i.e., below about yD=10, then
 3  
T yD
Kt ¼ K0 exp
yD bT

which is an exponential temperature dependence. The thermal conductivity should


therefore rise more rapidly than 1=T as the temperature is reduced in this region.
In fact, this exponential rise predicts an infinite conductivity at 0 K. Before this
happens, however, surface reflections of the phonons interfere, and the thermal con-
ductivity is held in check. As the absolute temperature approaches 0, the mean free

© 2005 by Marcel Dekker


5367-4 Flynn Ch05 R2 090904

Transport Properties of Solids 315

path increases until it reaches the order of the dimensions of the crystal. Phonons are
then reflected from its surfaces. With L and U now constant, Kt depends on the
dimensions of the crystal and varies as CV with temperature.
We have already seen that the heat capacity of a dielectric varies as T 3 at very
low temperatures, and so we also expect a T 3 temperature dependence for Kt at the
lowest temperatures. This dependence has in fact been observed experimentally.
The type of thermal conductivity curve suggested by all of these considerations
is exemplified by that of corundum in Fig. 5.4. Above temperatures corresponding to
about yD=10, phonon–phonon collisions govern. These are decreased at lower
temperatures to cause a maximum near yD=20, and below this the size of the crystal
limits the mean free path and the T 3 dependence causes a drop to zero at 0 K.
The phonon interaction and boundary scattering process establish the general
shape of the conductivity curve for a perfect crystalline dielectric. Further scattering
may occur at internal lattice imperfections such as vacancies, interstitial atoms,
impurity atoms (among which may be included isotopes of the primary constituents),
and dislocations. These added resistances have the effect of lowering and flattening
the conductivity curve and, in the limit of a glassy or disordered solid, lead to a curve
without a maximum.

1.2.2. Metallic Heat Conduction


Metals, having a lattice structure, conduct heat by phonons as do dielectrics. In
addition, heat may be conducted by the free electrons. These two mechanisms are
additive:

Kt ¼ Ke þ Kg ð5:18Þ

where Ke is the electron conduction; and Kg is the phonon conduction.


In metals having a large number of free electrons (e.g., Na, Ag, Au, Cu),
roughly one conduction electron per atom, Ke  Kg and Kg may be ignored.
Why do free electrons contribute so little to heat capacity but so much to
heat transport? Consider the following equation for electronic heat conduction,
analogous to the expression (Eq. (5.16)) for lattice heat conduction:

Ke ¼ ð1=3ÞCe Ue Le ð5:19Þ

where Ce is the electronic heat capacity; Ue the mean velocity of the electron and Le is
the mean free path of electrons between collisions with obstacles.
Although Ce is very small, Ue is the speed of light, not the speed of sound, as is
the case with phonons. Furthermore, electrons effectively scatter phonons, reducing
the contribution of Kg still further. In fact, the striking difference between the very
high phonon conduction of dielectrics and its very low value in metals is due to
the effectiveness of the free electrons in scattering phonons. Accordingly, Ke is much
greater than Kg for most pure metals.
The details of thermal conductivity in metals can be seen by examining the
temperature dependence of the electron heat capacity Ce and the lattice heat capacity
Cl. The former is given by

k2
Ce ¼ p2 ZNo T ð5:20Þ
m v2

© 2005 by Marcel Dekker


5367-4 Flynn Ch05 R2 090904

316 Chapter 5

where m is the effective electron mass: n is the electron velocity at the energy of the
Fermi face: Z the number of free electrons per atom and No is Avogadro’s number
(number of atoms per mole). The coefficient of T in the preceding equation is a
constant of the material, and it is convenient to give it the symbol g. Then

Ce ¼ gT ð5:21Þ

that is, the electron specific heat varies linearly with T. Ordinarily, gT is small com-
pared to the lattice specific heat. (See Sec. 1.1.1.) However, it becomes appreciable at
very high T, where the lattice CV remains constant at a value 3R. The electron CV
also becomes appreciable at low T, where the Debye model predicts that the lattice
CV varies as (T=yD)3. Then the total CV is given by
 3
T
CV ¼ gT þ 234R ð5:22Þ
yD
For copper, at 1 K, the first term is roughly four times the magnitude of the
second term in this equation.
In dealing with the thermal conductivity due to the electrons, CV0 , the specific
heat per unit volume, is needed. This differs from CV, the specific heat per mole,
by having No replaced by n, the number of electrons per unit volume.
For metals, the thermal conductivity is almost entirely the result of energy
carried by the electrons rather than by the phonons. The two heat fluxes are additive,
but the phonon conductivity becomes negligible because collisions of phonons with
electrons drastically reduce the phonon mean free path in metals.
In treating the thermal conductivity due to electrons, the solid is considered to
contain an electron gas. The gas thermal conductivity expression may again be
applied, that is

K ¼ ð1=3ÞCV0 UL ð5:23Þ

The manner in which K varies with T for an electrical conductor is indicated in


Fig. 5.5a. The causes of this variation may be seen by looking at how the terms of
Eq. (5.23 ) vary with temperature. C0 V ¼ g T is plotted in Fig. 5.5b. The electron velo-
city V is independent of T as indicated in Fig. 5.5c. The mean free path varies with T

Figure 5.5 Thermal conductivity by electron mechanisms.

© 2005 by Marcel Dekker


5367-4 Flynn Ch05 R2 090904

Transport Properties of Solids 317

in much the same way as for phonons. At high T, the electrons collide with the
phonons, so L decreases as the number of phonons increases. At low temperature,
the number of phonons is small and the electron mean free path is limited by dis-
tances between crystal imperfections. These are impurities, dislocations, intercrystal-
line boundaries, or, finally, the specimen walls. Looking again at the product of CV0 ,
V, and L, the quantity V is constant, so K depends on CV0 and L. At low T, L is
constant, so K follows CV0 and varies linearly with T. At high T, CV0 still varies linearly
with T, but L begins to fall off so rapidly that the product begins to drop off. At
intermediate T, a broad maximum occurs.
Figures 5.6 and 5.7 show the thermal conductivity curves for several solids,
with the highest curve having the highest purity. It will be noted that the metals
of high purity exhibit a maximum of conductivity at low temperatures that in some
cases is many times the room temperature value. Moreover, the conductivities of
these metals approach a room temperature value that is almost temperature-
independent. The height (or existence) of the maximum is quite sensitive to certain
slight impurities, although it will be noted that some of the commercial coppers
exhibit maxima.
Table 5.3 gives the thermal conductivity and thermal conductivity integral for a
few common materials of construction (Johnson, 1960).
The variability of thermal conductivity with alloying and cold working and the
suppression of the maximum are shown clearly in Figs. 5.5 and 5.6. Inconel, Monel,

Figure 5.6 Thermal conductivity of several coppers.

© 2005 by Marcel Dekker


5367-4 Flynn Ch05 R2 090904

318 Chapter 5

Figure 5.7 Thermal conductivity of some common solids.

and stainless steel are structural alloys that also exhibit these properties and are thus
useful in cryogenic service, which requires low thermal conductivity over the entire
temperature range.
Alloys and highly cold-worked metals represent cases in which scattering from
imperfections in the lattice is so great that this effect predominates over phonon
scattering even at room temperature. The resulting flattening of the thermal conduc-
tivity curve is illustrated in Fig. 5.4 for nickel–chromium stainless steel and brass.
Furthermore, pure metals having a small number of conduction electrons per atom
(e.g., antimony and bismuth) and semiconductors (e.g., germanium) are all metals
for which phonon thermal conduction can no longer be safely neglected.
Experimental thermal conductivity values for most materials used in
low-temperature design are readily available in the literature (Powell, 1963;
Touloukian, 1964; Powell et al., 1966).
Example 5.3. In support members for cryogenic storage vessels and other equip-
ment, it is desirable to minimize the heat transfer through the members but maximize

© 2005 by Marcel Dekker


Transport Properties of Solids
Table 5.3 Thermal Conductivities and Thermal Conductivity Integrals for Several Materialsa

Cu (beryllium Aluminium Low-carbon Stainless Monel


tough pitch) (6063 T5) steel (C1020) steel (304) (drawn) Teflon

Temperature RT RT RT RT RT RT
(K) Kt 4:2 K Kt dT Kt 4:2 K Kt dT Kt 4:2 K Kt dT Kt 4:2 K Kt dT Kt 4:2 K Kt dT Kt 4:2 K Kt dT

4.2 1.9 0 34 0 3.0 0 0.24 0 0.43 0 0.046 0.0


10 4.8 19 86 360 11.5 43 0.77 2.9 1.74 6.3 0.096 0.44

5367-4 Flynn Ch05 R2 090904


20 10.6 95 170 1650 24.0 222 1.95 16.3 4.30 36.4 0.141 1.64
30 16.2 229 230 3650 32.0 502 3.30 42.4 6.90 92.9 0.174 3.23
40 21.0 415 270 6200 38.6 867 4.70 82.4 9.00 173 0.193 5.08
50 26.1 650 280 8950 47.6 1310 5.80 135 10.95 273 0.208 7.16
60 30.0 930 270 11700 53.6 1810 6.80 198 12.09 368 0.219 9.36
70 33.7 1250 248 14300 57.5 2360 7.60 270 13.06 513 0.228 11.6
80 37.0 1600 230 16700 60.0 2950 8.26 349 13.90 647 0.235 13.9
90 40.1 1990 222 19000 61.8 3350 8.86 436 14.63 791 0.241 16.3
100 43.0 2400 216 21100 62.9 4170 9.40 528 15.27 940 0.245 18.7
120 48.4 3300 207 25300 64.1 5450 10.36 726 16.26 1260 0.251 23.7
140 53.3 4320 201 29300 64.6 6750 11.17 939 17.34 1590 0.255 28.7
160 57.6 5440 200 33300 64.8 8050 11.86 1170 18.25 1950 0.257 33.8
180 61.5 6640 200 37300 64.9 9350 12.47 1410 19.02 2320 0.258 39.0
200 65.0 7910 200 41300 65.0 10700 13.00 1660 19.69 2710 0.259 44.2
250 72.4 11300 200 51300 65.0 13900 14.07 2340 21.02 3730 0.260 57.2
300 78.5 15000 200 61300 65.0 17200 14.90 3060 22.00 4800 0.260 70.2
a
RT
The thermal conductivity (Kt) is expressed in dimensions of W=(m K), and the thermal conductivity integral ( 4:2 K Kt dT) is given in dimensions of W=m.

319
© 2005 by Marcel Dekker
5367-4 Flynn Ch05 R2 090904

320 Chapter 5

the strength of the members at the same time. An important parameter in this situa-
tion is the strength-to-conductivity ratio. Determine this ratio, Sy=Kt [in (N K)=(W
m)] at 22.2 K for
a. 2024 T4 aluminum
b. 304 stainless steel
c. Monel
d. Beryllium copper
e. Teflon

Using Fig. 4.21 for Sy values, and Fig. 5.7 for Kt values:
a. 2024 T4 aluminum
Sy ¼ 7.24  108 N=m2
Kt ¼ 17.3 W=(m K)
Sy=Kt ¼ 4.18  107 N K=(W m)
b. 304 stainless steel
Sy ¼ 1.32  109 N=m2
Kt ¼ 2.77 W=(m K)
Sy=Kt ¼ 4.77  108 N K=(W m)
c. Monel
Sy ¼ 1.03  109 N=m2
Kt ¼ 7.61 W=(m K)
Sy=Kt ¼ 1.35  108 N K=(W m)
d. Beryllium copper
Sy ¼ 8.96  108 N=m2
Kt ¼ 9.52 W=(m K)
Sy=Kt ¼ 9.41  107 N K=(W m)
e. Teflon
Sy ¼ 1.38  108 N=m2
Kt ¼ 0.138 W=(m K)
Sy=Kt ¼ 1.0  109 N K=(W m)

1.2.3. Design Considerations


It is difficult to predict thermal conductivities of nearly pure metals on the basis of
their chemical analyses because (1) the effect of each kind of impurity is specific and
depends on its electronic band structure and (2) the effect of a given impurity is much
greater if it is in solid solution than if it is segregated at grain boundaries. Thus a
gross or overall chemical analysis provides insufficient information as shown by the
bands of Fig. 5.8.
It is believed that the relatively poor thermal conductivity of OFHC (oxygen-
free, high-conductivity) copper is due to the latter effect (solution vs. inclusion). The
removal of the oxygen may have permitted metallic impurities that were previously
segregated as oxides at the grain boundaries to enter the lattice. On the other hand,
the good conductivities of the free-machining lead and tellurium coppers may exist in
spite of the additives because the impurities in this case are soluble only slightly (if at

© 2005 by Marcel Dekker


5367-4 Flynn Ch05 R2 090904

Transport Properties of Solids 321

Figure 5.8 Thermal conductivity of selected classes of materials.

all) in the copper. Thus they are localized at grain boundaries. This condition is also
responsible for the free-machining characteristic.
The extreme sensitivity of the low-temperature conductivity of nearly pure
metals to impurities and cold working is in strong contrast to the specific heat and
expansivity, both of which are somewhat sensitive to the type of lattice structure
and to the interatomic forces but relatively insensitive to local imperfections of the
lattice. The variability shown by coppers is demonstrated by other metals. This
serves to illustrate that where conductivity is concerned, there are no ‘‘pure’’ metals
but only alloys of various degrees of dilution. Fortunately for the cryogenics
designer, the only other widely used cryogenic metal (aluminum) is available in only
one nominally pure commercial grade.
As noted earlier, the lattice conduction of pure metals and dilute metal alloys is
generally small, so ke can be taken as the total conduction. This principle has been
used as the basis for a generalized low temperature equation for thermal conductivity
as represented in Fig. 5.9, where k is equal to k=km, T  is equal to T=Tm, and Tm is
the temperature at which k is a maximum, km. In analogy to reduced equations of
state for thermodynamic properties of fluids, this correlation is referred to as the
principle of corresponding states. The reducing parameters chosen are the tempera-
ture and thermal conductivity at the maximum in conductivity. The results of this
correlation are illustrated in Fig. 5.9. The figure represents 1002 data points for a
total of 22 metals (83 specimens) in varying states of impurity and imperfection
content. The standard deviation of the data is only 3.2%.

1.3. Thermal Expansivity


In our discussion of thermal properties so far, it has been assumed that the atoms of
a solid vibrate symmetrically about some equilibrium position in the solid. If this

© 2005 by Marcel Dekker


5367-4 Flynn Ch05 R2 090904

322 Chapter 5

Figure 5.9 Reduced thermal conductivity vs. reduced temperature of 22 metallic elements.
The k is k=km where km is the maximum value of k and T is T=Tm where Tm is the tempera-
ture corresponding to km.

were the case, there would be no thermal expansion. Instead, the intermolecular
potential energy curve (forces of attraction and repulsion) is not symmetrical.
The general shape of the intermolecular potential, as illustrated in Fig. 5.10, is
asymmetric. This results from a repulsive force that increases very rapidly as the
two atoms approach each other closely (one atom is considered at rest at the origin
in Fig. 5.10) plus an attractive force that is not such a strong function of distance.
Their algebraic sum provides a curve having a minimum at r0, the classical
equilibrium separation of the atoms at 0 K. Increasing the temperature from 0 K
corresponds to raising the energy of the system slightly, and now thermal vibrations
take place between two higher values of r. But because of the asymmetry of the
potential well, the average distance no longer corresponds to r0; it is slightly greater
and results in the observed thermal expansion of materials. The mean spacing of the
atoms increases with temperature as the energy (temperature) of the material
increases. Thus, the coefficient of thermal expansion increases as temperature is
increased.
Both the specific heat and the coefficient of thermal expansion arise from the
intermolecular potential, and accordingly these two properties are related. Our
present understanding of this relation is largely due to Gruneisen (1926), who
developed an equation of state for solids based on the lattice dynamics of Debye.
He showed that the dimensionless ratio g should be a universal constant:

g ¼ aV =bT CV ð5:24Þ

© 2005 by Marcel Dekker


5367-4 Flynn Ch05 R2 090904

Transport Properties of Solids 323

Figure 5.10 Intermolecular potential curve.

or

g ¼ aV =bS CV ð5:25Þ

where
 
1 @V
a¼ volume coefficient of thermal expansion
V @T P
 
1 @V
bT ¼  isothermal compressibility
V @P T
 
1 @V
bS ¼  adiabatic compressibility
V @T S

Here the volumetric coefficient of thermal expansion a is defined as the


fractional change in volume per unit change in temperature as the pressure on the
material remains constant.
The linear coefficient of thermal expansion lt is defined as the fractional change
in length (or any linear dimension) per unit change in temperature while the stress on
the material remains constant. For isotropic materials, a ¼ 3lt.
The Gruneisen relation can also be expressed in terms of the more commonly
measured mechanical properties as
gCV r
a¼ ð5:26Þ
B
where r is the density of the material; B the bulk modulus (see Eq. (4.1)); g the
Gruneisen constant and CV is the specific heat and
a ¼ ð1=3Þlt for isotropic materials
where lt ¼ linear coefficient of thermal expansion.

© 2005 by Marcel Dekker


5367-4 Flynn Ch05 R2 090904

324 Chapter 5

Since both CV and a depend on the intermolecular potential of the lattice, they
are related to a first approximation by the Debye temperature yD

V dyD
g¼ ð5:27Þ
yD dV

where V is the volume of the material.


Some values of the Gruneisen constant computed from room temperature
property data are given in Table 5.4. In general, the Gruneisen constant has a value
close to 1.5 for the bcc metals, about 2.3 for the fcc metals, and in all cases it tends to
increase with increasing atomic number.
For most materials, volume and compressibility are both approximately tem-
perature-independent, and since both vary in the same direction with temperature
their ratio is even more constant. Hence if g were temperature-independent, one
could hold that a should vary in proportion to the specific heat and should therefore
have a temperature variation like the CV (Debye) curve shown earlier. This is indeed
the case and provides a useful correlation for thermal expansion data. In fact, at low
temperatures (T < yD=12), the coefficient of thermal expansion is proportional to
T 3, in accord with the Debye expression.
Example 5.4. Determine the Gruneisen constant for beryllium copper at 22.2 K if
copper has a density of 8304 kg=m3 and a Poisson ratio of 0.335. Assume that the
Debye expression is valid at this temperature

yD ¼ 309; MW ¼ 63:54
T=yD ¼ 22=309 ¼ 0:0712 < 1=12
 3
T
CV ¼ 233:8R
yD
 
1:987
CV ¼ ð233:8Þ ð0:0712Þ3 ¼ 0:00264 cal=ðg KÞ ¼ 11:04 J=ðkg KÞ
63:54

Table 5.4 Values of the Gruneisen Constant of Selected Cryogenic


Materials Computed from Room Temperature Properties

Material g Material g

Li 1.17 Mo 1.57
Na 1.25 W 1.62
K 1.34 Fe 1.60
Mg 1.51 Ni 1.88
Al 2.17 Co 1.87
Sn 2.14 Pd 2.23
Pb 2.73 Pt 2.54
Sb 0.92 NaCl 1.63
Bi 1.14 KCl 1.60
Cu 1.96 KBr 1.68
Au 3.03 FeS2 1.47
Zn 2.01 PbS 1.94
Cd 2.19 Ta 1.75

© 2005 by Marcel Dekker


5367-4 Flynn Ch05 R2 090904

Transport Properties of Solids 325

For isotropic materials, b ¼ 3lt

lt ¼ 0:1  106 = R ¼ 1:8  107 =K


b ¼ 5:4  107 =K
E

3ð1  2mÞ
E ¼ 1:31  1011 N=m3
m ¼ 0:335
1:31  1011 N=m2
B¼ ¼ 1:32  1011 N=m2
3½1  2ð0:225Þ
bB ð5:4  107 K1 Þð1:32  1011 N=m2 Þ
g¼ ¼
CV r ½11:04 J=ðkg KÞ ð8304 kg=m3 Þ
g ¼ 0:78
Figures 5.11 and 5.12 illustrate the expansivity of several metals as a function
of temperature and show that the behavior predicted by Eq. (5.26) is actually
observed experimentally. Namely, the coefficient of thermal expansion b shows
the same temperature dependence as that of the heat capacity CV. Two sources
(Corruccini and Gniewek, 1961; White, 1959) may be consulted for tabulations of
the mean linear thermal expansion of other selected construction materials.
In general, organic polymers have expansivities that are considerably larger
than those of metals. Expansivities of such materials may be reduced considerably

Figure 5.11 Coefficient of linear expansion for several metals as a function of temperature
in K.

© 2005 by Marcel Dekker


5367-4 Flynn Ch05 R2 090904

326 Chapter 5

Figure 5.12 Thermal expansion of steels vs. temperature in  R.

by the addition of filler material of low expansivity such as glass fiber, silica,
alumina, and asbestos.
As in the case of heat capacity, anomalies can also occur in the thermal expan-
sivity of materials. Various transformations that occur in the solid state can alter the
interatomic forces of the material and lead to changes in its overall dimensions. For
example, the ferromagnetic Curie point of dysprosium at low temperatures leads to
an irregularity in its expansivity. Other examples are KH2PO4 with a ferroelectric
Curie point at 122 K, and the ‘‘glass’’ transitions of soft rubbers. In fact, the mea-
surement of thermal expansion is a useful method by which solid-state transitions
can be investigated. Tables 5.5–5.7 give values of thermal expansion for some
structural materials at low temperatures.
Example 5.5. An aluminum distillation column 20 m high is placed in an air separa-
tion service. How much will the column shrink as it is cooled from 300 K to an aver-
age temperature of 89 K?

a at 89 K ¼ 35  105 ; a at 300 K ¼ 431  105


From 89 to 300 K
DL
a ¼ ð431  35Þ  105 ¼ 396  105 ;
 a ¼
L0
or
a ¼ ð20 mÞð396  105 Þ ¼ 7:92  102 m
DL ¼ L0 
The third law of thermodynamics predicts that the expansivity of all materials
must go to 0 at 0 K. At very low temperatures, a few materials show an anomalous
expansion with decreasing temperature. In addition to 304 stainless steel and Pyrex
glass, uranium, fused silica, and the intermetallic compound indium antimonide have

© 2005 by Marcel Dekker


5367-4 Flynn Ch05 R2 090904

Transport Properties of Solids 327

Table 5.5 Mean Linear Thermal Expansion of Various Metals at Selected Temperaturesa

1020
Tempera- Low-carbon Stainless Yellow
ture (K) Ni Mg Zn Cu Al steel steel Inconel Monel brass

0 0 0 0 0 0 0 0 0 0 0
20 0 1 1 1 0 0 0 0 0 1
40 1 5 9 4 3 1 0 1 1 4
60 4 12 28 12 9 4 4 5 6 16
80 12 29 57 26 24 10 13 12 15 34
100 23 55 93 45 46 20 32 24 28 58
120 38 87 133 67 71 32 49 38 45 85
140 55 124 176 92 104 47 72 55 64 115
160 74 164 221 120 138 63 97 74 84 147
180 95 208 267 149 175 81 122 95 107 180
200 117 254 314 179 214 101 150 117 130 215
220 140 303 363 210 254 121 179 140 154 249
240 164 353 413 241 295 142 208 163 179 283
260 188 403 465 273 339 164 240 187 207 320
280 213 453 518 305 384 187 272 212 233 357
300 239 503 572 337 418 210 304 238 261 397
a
Net expansion from 0 K to any other temperature in the table, DL=L0  105. DL is the change in length
from 0 K to T. L0 is the length at 0 K.
Source: Corruccini and Gniewek (1961).

been reported to exhibit this characteristic. (This effect for Pyrex glass between 40 C
and 0 C can be seen in Table 5.6.) Some anisotropic materials such as zinc show a
negative expansivity in one crystallographic direction although they have an average
positive expansivity, but for the materials just mentioned, it is the average linear
expansivity that has been reported negative. In 304 stainless steel, this occurrence
is ascribed to the partial transformation of the material at low temperatures to a sec-
ond phase, martensite.
Figure 5.13 shows the range of Thermal expansion values for various
categories of materials.

2. EMISSIVITY, ABSORPTIVITY, AND REFLECTIVITY

Estimating radiant heat transfer across an insulating vacuum requires data on the
emissivity, absorptivity, and reflectivity of the surfaces involved. The common defini-
tions for these terms are as follows.
Total normal emissivity (En) is the rate of radiant energy emission normal to the
surface, divided by the corresponding rate from a blackbody.
Total hemispherical emissivity (eh) is the rate of radiant energy emission into a
hemisphere (centered on the normal to the emitting surface), divided by
the corresponding rate from a blackbody.
Total absorptivity (a) is the fraction of energy incident upon a surface that is
absorbed. Quantities an and ah analogous to en and eh may be distin-
guished.
Reflectivity (r) is the fraction of incident energy that is reflected; r ¼ 1  a. The
value of the reflectivity also depends on the angle of incidence.

© 2005 by Marcel Dekker


328
Table 5.6 Mean Linear Thermal Expansion of Various Nonmetals at Selected Temperaturesa

Molded polyester Cast Cast


Temperature rod reinforced phenolic epoxy Nylon
(K) Pyrex Teflon with glass fiber rod polymer rod Fluoroethylene Polystyrene

0 0 0 0 0 0 0 0 0
20 1 30 3 14 10 10 21 27

5367-4 Flynn Ch05 R2 090904


40 2 90 11 38 39 37 65 82
60 1.5 160 21 70 78 81 116 152
80 þ1 245 34 109 126 142 173 235
100 4.5 335 49 154 181 217 235 329
120 8.5 430 67 205 242 301 301 432
140 13 540 88 261 310 393 372 542
160 17.5 660 110 321 385 493 444 658
180 22.5 805 134 385 467 600 531 778
200 27.5 995 159 452 556 716 618 900
220 33 1215 184 524 651 841 711 1024
240 39 1440 210 602 753 977 811 1152
260 44.5 1670 237 688 862 1124 921 1284
280 50.5 1900 264 782 980 1282 1045 1422
300 57 2460 291 889 1107 1450 1187 1566
a
Net expansion from 0 K to any other temperature in the table, DL=L0  105.
Source: Corruccini and Gniewek (1961).

Chapter 5
© 2005 by Marcel Dekker
Transport Properties of Solids
Table 5.7 Thermal Expansion of Selected Metalsa

Aluminum Aluminum Aluminum Titanium Copper ETP Brass Monel K Stainless 2800 Steel
2024 5083 7039 6A1-4V & OFHC 70=30 &S steel 304 (9% Ni)
Temperature
( R) A B A B A B A B A B A B A B A B A B

5367-4 Flynn Ch05 R2 090904


0 434 0 455 0 462 0 187 0 356 0 410 0 277 0 326 0 232 0
20 434 0.03 455 0.03 462 0.05 187 0.01 356 0.03 410 0.07 277 0.02 326 0.02 232 0.02
40 434 0.27 455 0.27 462 0.32 187 0.08 356 0.23 409 0.40 277 0.14 326 0.15 232 0.09
60 433 0.78 454 0.80 461 0.95 187 0.24 355 0.73 408 1.31 276 0.44 325 0.40 231 0.29
80 431 1.58 452 1.63 458 1.82 186 0.51 353 1.62 404 2.47 275 0.98 324 0.85 230 0.62
100 427 2.56 447 2.63 453 3.02 184 1.00 348 2.64 398 3.59 272 1.61 322 1.83 229 1.08
120 420 3.65 441 3.74 446 4.05 182 1.56 342 3.61 389 4.60 268 2.33 317 3.05 226 1.60
140 412 4.70 432 4.82 438 5.10 178 2.01 334 4.50 379 5.56 263 3.07 310 4.00 222 2.13
160 402 5.68 422 5.82 426 6.06 174 2.45 324 5.21 367 6.36 256 3.66 301 4.78 218 2.62
180 389 6.55 409 6.72 413 6.89 168 2.81 313 5.84 354 7.06 248 4.16 291 5.33 212 3.07
200 376 7.26 395 7.50 398 7.60 162 3.09 301 6.33 339 7.54 239 4.63 280 5.75 205 3.50
220 360 7.94 379 8.18 382 8.26 156 3.37 288 6.74 324 7.93 230 5.00 268 6.13 198 3.92
240 344 8.47 362 8.77 365 8.81 149 3.61 274 7.13 308 8.21 219 5.33 255 6.48 190 4.32
260 327 8.95 344 9.29 347 9.30 142 3.82 259 7.46 291 8.48 208 5.62 242 6.78 181 4.67
280 308 9.38 325 9.70 328 9.73 134 4.00 244 7.75 274 8.74 197 5.89 228 7.02 171 4.99
300 289 9.74 305 10.10 308 10.13 126 4.15 228 7.96 256 8.96 185 6.14 214 7.24 161 5.24
320 269 10.03 285 10.49 288 10.49 117 4.28 212 8.15 238 9.14 172 6.34 199 7.43 150 5.47
340 249 10.30 264 10.80 266 10.80 108 4.42 196 8.30 219 9.31 159 6.51 184 7.59 139 5.65

329
(Continued)

© 2005 by Marcel Dekker


Table 5.7 (Continued )

330
Aluminum Aluminum Aluminum Titanium Copper ETP Brass Monel K Stainless 2800 Steel
2024 5083 7039 6A1-4V & OFHC 70=30 &S steel 304 (9% Ni)
Temperature
( R) A B A B A B A B A B A B A B A B A B

360 228 10.56 242 11.06 244 11.10 99 4.55 179 8.44 201 9.46 146 6.67 169 7.78 128 5.81
380 207 10.79 219 11.31 222 11.37 90 4.64 162 8.56 182 9.60 133 6.82 153 7.92 116 5.96
400 185 10.98 196 11.55 199 11.62 81 4.73 145 8.68 162 9.72 119 6.97 137 8.07 104 6.09
420 163 11.15 173 11.75 175 11.85 71 4.81 127 8.78 143 9.84 105 7.10 121 8.19 92 6.21
440 141 11.30 149 11.94 151 12.06 61 4.89 110 8.87 123 9.95 91 7.22 104 8.32 79 6.32
460 118 11.47 125 12.13 127 12.27 51 4.97 92 8.96 103 10.05 76 7.33 87 8.45 66 6.42

5367-4 Flynn Ch05 R2 090904


480 95 11.60 101 12.31 102 12.45 41 5.04 74 9.05 83 10.15 61 7.44 70 8.56 53 6.51
500 72 11.73 76 12.47 77 12.62 31 5.11 56 9.15 62 10.25 46 7.54 53 8.67 40 6.60
520 48 11.85 51 12.62 52 12.79 21 5.17 38 9.25 42 10.3 31 7.63 36 8.78 27 6.68
540 24 11.97 26 12.77 26 12.96 11 5.23 19 9.33 21 10.4 16 7.72 18 8.89 14 6.76
560 0 12.08 0 12.90 0 13.12 0 5.28 0 9.40 0 10.5 0 7.82 0 9.00 0 6.84
a
A is the net linear contraction coefficient, the change in length from the base temperature of 560 R to any temperature in the table. A  105 ¼ (L560  L)=L560, in m=m, in.=in.,
etc. B is the coefficient of linear expansion at the temperature shown. B  106 ¼ (DL=DT )(1=L560), per degree Rankine in m=(m  R) in.=(in.  R). etc.
Source: Schmidt (1968).

Chapter 5
© 2005 by Marcel Dekker
5367-4 Flynn Ch05 R2 090904

Transport Properties of Solids 331

Figure 5.13 Range of thermal expansion values for various categories of materials.

The rate of energy emission from a blackbody at temperature T is


Q ¼ s T4
where s is the Stefan–Boltzmann blackbody radiation constant, and T is the Kelvin
temperature.
The emissivities of electrical conductors are obtained theoretically from
electromagnetic theory provided the surfaces are smooth and oxide-free and the
wavelength of radiation is not too small. For these types of pure materials, the
emissivity is directly proportional to the absolute temperature of the material
T=51:3
n¼ ð5:28Þ
ke;4921=2
where ke,492 is the electrical conductivity at 492 R in mho=cm and T is the absolute
temperature in degrees Rankine.
Equation (5.28) shows that the emissivity of a pure electrical conductor is
directly proportional to its absolute temperature. Emissivity is also shown to be
related to the electrical conductivity for electrical conductors, as both are electro-
magnetic in origin.
For the same reason, the emissivity of electrical insulators can be related to the
index of refraction nr of the material through electromagnetic theory. For electrical
insulators,
4nr
E¼ ð5:29Þ
ðnr þ 1Þ2
for normal incidence of radiation. Since the index of refraction for most electrical
insulators is less than 2, the emissivity of insulators is near unity. [For nr ¼ 2,
E ¼ 8=9 ¼ 0.889 from Eq. (5.29).]

© 2005 by Marcel Dekker


5367-4 Flynn Ch05 R2 090904

332 Chapter 5

For electrical conductors, it is generally true that any process, such as


annealing, that increases the electrical conductivity of a material also decreases its
emissivity. In addition, the following approximate generalizations can be drawn
from experimental observations:
1. Materials having the lowest emissivities also have the lowest electrical
resistances.
2. Emissivity decreases with decreasing temperature.
3. The apparent emissivity of good reflectors is increased by surface contami-
nation.
4. Alloying a metal increases its emissivity.
5. Emissivity is increased by treatments such as mechanical polishing that
result in work hardening of the surface layer of the metal.

At a fixed temperature, the emissivity must equal the absorptivity. (If these
differed, there could be a net transfer of heat between two surfaces at the same
temperature—a violation of the second law of thermodynamics.) Helpful tabulations
of emissivity values have been prepared by Fulk (1959), and the emissivities of a few
selected materials are shown in Table 5.8.

3. ELECTRICAL PROPERTIES

In metals, the primary mechanisms of thermal conductivity and electrical conductiv-


ity are the same. Accordingly, the mechanisms of electron–lattice interaction
described in Sec. 1 for thermal conductivity apply equally well to electrical conduc-
tivity. This fact becomes the starting point for the present discussion.
Superconductivity, to be discussed in Sec. 4, also depends on the electron–
lattice interaction, albeit in a much more subtle way.

3.1. Electrical Resistivity and Conductivity


3.1.1. Metals
The most striking property of metals, in fact the one used to define this class of mate-
rials, is electrical conductivity.
An electrical field applied to a metal produces a drift of the free electrons. The
resistance that this flow encounters is due to the scattering of electrons by obstacles
in their path. This scattering is provided (1) by the thermal vibrations of the crystal
lattice, i.e., the phonons, and (2) by lattice imperfections, which we will consider in
this order. The effect of temperature on resistance can be deduced by considering the
effect of temperature on these two resistance mechanisms, that is, the collision of
electrons with phonons and with imperfections in the crystal lattice. The first is a
dynamic scattering mechanism; the second, static. As in the case of thermal conduc-
tivity, these scattering mechanisms can be considered independent to a very good
approximation, and we can write for the electrical resistivity, r,
r ¼ ri þ rr ð5:30Þ
where ri is the phonon scattering resistivity, the resistivity of an ideal crystal; rr
is the imperfection scattering, the ‘‘residual’’ resistivity of a real crystal at 0 K, when
phonon scattering is inoperative. This is Matthiessen’s rule, observed empirically on
metals of different purity.

© 2005 by Marcel Dekker


5367-4 Flynn Ch05 R2 090904

Transport Properties of Solids 333

T