You are on page 1of 146

Catalog No.

L40000e

Materials of Construction for Use in an LNG Pipeline

CONTRACT PR-3-42

Prepared for the


Materials Committee
Pipeline Research Committee

of
Pipeline Research Council International, Inc.

Prepared by the following Research Agencies:


BATTELLE

Authors:
J. Dainora
A.R. Duffy
T.J. Atterbury

Publication Date:
April 1968
“This report is furnished to Pipeline Research Council International, Inc. (PRCI) under
the terms of PRCI NN-18 Report No. 111, between PRCI and BATTELLE. The contents
of this report are published as received from BATTELLE. The opinions, findings, and
conclusions expressed in the report are those of the authors and not necessarily those
of PRCI, its member companies, or their representatives. Publication and dissemination
of this report by PRCI should not be considered an endorsement by PRCI or BATTELLE,
or the accuracy or validity of any opinions, findings, or conclusions expressed herein.

In publishing this report, PRCI makes no warranty or representation, expressed or


implied, with respect to the accuracy, completeness, usefulness, or fitness for purpose of
the information contained herein, or that the use of any information, method, process, or
apparatus disclosed in this report may not infringe on privately owned rights. PRCI
assumes no liability with respect to the use of, or for damages resulting from the use of,
any information, method, process, or apparatus disclosed in this report.

The text of this publication, or any part thereof, may not be reproduced or transmitted in
any form by any means, electronic or mechanical, including photocopying, recording,
storage in an information retrieval system, or otherwise, without the prior, written
approval of PRCI.”

Pipeline Research Council International Catalog No. L40000e

Copyright, 1968
All Rights Reserved by Pipeline Research Council International, Inc.

PRCI Reports are Published by Technical Toolboxes, Inc.

3801 Kirby Drive, Suite 340


Houston, Texas 77098
Tel: 713-630-0505
Fax: 713-630-0560
Email: info@ttoolboxes.com
PRCI Pipeline Research Committee

Chairmen

W. A. Haas, Northern Natural Gas Company, Omaha, Nebraska (1966-1967)


A. J. Shoup, Texas Eastern Transmission Corporation, Houston, Texas (1968-)

Vice Chairmen

A. J. Shoup, Texas Eastern Transmission Corporation, Houston, Texas (1966-1967)


R. D. Morel, Algonquin Gas Transmission Company, Boston, Massachusetts (1968-)

Members

S. A. Bradfield, Southern California Gas Company, Los Angeles, California


O. W. Clark, Southern Natural Gas Company, Birmingham, Alabama
R. H. Crowe, Transcontinental Gas Pipe Line Corporation, Houston, Texas
O. C. Davis, Natural Gas Pipeline Company of America, Chicago, Illinois
J. F. Eichelmann, El Paso Natural Gas Company, El Paso, Texas
J. L. Gere, Cities Service Gas Company, Oklahoma City, Oklahoma
L. E. Hanna, Panhandle Eastern Pipe Line Company, Omaha, Nebraska
G. W. McKinley, Con-Gas Service Corporation, Pittsburgh, Pennsylvania
L. D. Myers, United Gas Pipeline Company, Shreveport, Louisiana
S. Orlofsky, Columbia Gas System Service Corporation, New York City
R. R. Olson, Colorado Interstate Gas Company, Colorado Springs, Colorado
J. L. Parrish, Tennessee Gas Pipeline Company, Houston, Texas
A. W. Stanzel, Michigan Wisconsin Pipe Line Company, Detroit, Michigan
H. L. Stowers, Texas Gas Transmission Corporation, Owensboro, Kentucky
T. E. Walsh,Pipeline Research Council International, Inc., Inc., New York City

PRCI Supervising Committee for Project PR-3-42

Chairmen

C. E. Schorre, Transcontinental Gas Pipe Line Corporation, Houston,Texas (1966-1967)


W. H. Penn, Tennessee Gas Pipeline Company, Houston, Texas (1968-)

Members

B. Clark, Northern Natural Gas Company, Omaha, Nebraska


S. J. Cunningham, Southern California Gas Company, Los Angeles, California
L. L. Elder, Columbia Gas System Service Corporation, Columbus, Ohio
L. L. Stallard, Natural Gas Pipeline Company of America, Chicago, Illinois
T. E. Walsh, Pipeline Research Council International Inc., New York, New York

Printed in the United States of America


Copyright 1968, by Pipeline Research Council International, Inc.
April, 1968 Price $7.00
This Page Intentionally Left Blank
FOREWORD

The recent activities in the field of liquefied natural gas


(LNG) has induced considerable interest in the possibility of piping
natural gas in its liquid, rather than gaseous, form. Many national and
international LNG projects are already commercially established, These
plants are presently piping LNG within the plant and to and from loading
terminals. Expansion of these facilities including piping LNG between
widely separated storage depots is an anticipated future step in the
advance of the LNG industry.

In 1966, the Pipeline Research Committee of the Pipeline Research Council


International, Inc. began Project PR-3-42 to investigate materials of construction
for LNG pipelines. The objective of the project during the first year was
to survey commercially available materials considered currently acceptable
from a technical viewpoint for the construction of an LNG pipeline.

The survey has indicated that construction materials exist such


that transmission of LNG over considerable distances does appear to be
t e c h n i c a l l y f e a s i b l e . Although the study has identified pipe and insulation
materials which are suitable for some applications, it is not the intention
to sanction or express disapproval of specific products. The report merely
reflects available published data.

Because materials of construction are somewhat dependent upon


the design configuration and the transmission distance, the report includes
a basic treatment of these factors. While all of the problems of design and
operation associated with LNG piping have not been investigated, it appears
that design configurations are possible to cope with at least the major
problems, i.e., thermal contraction of the line and heat-influx to the LNG.

The results of this study are being published in this report


because they are believed to be an important source of information for the
natural gas industry and a significant contribution to the state-of-the-art
in the design of cryogenic pipelines.

W. H. Penn, Chairman
Supervisory Committee
for Project PR-3-42
This Page Intentionally Left Blank
TABLE OF CONTENTS
Page

INTRODUCTION . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1

SUMMARY . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3
MODULAR DISTANCE STUDIES . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8

Mathematical Analysis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8
Calculations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9
Heat-Transfer Resistance . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9
Pumping- and Cooling-Station Spacing . . . . . . . . . . . . . . . . . . . . . . . . . . 10
Horsepower Requirements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20
DESIGN CONSIDERATIONS . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 27

Fully Restrained Line . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 28


Expansion Loops . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 32
Expansion Joints ....................... ..... ..................... 34
Allowable Design Stress of Pipe . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 35
PIPE MATERIALS . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 38
Significant Properties . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 38
Resistance to Fracture . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 39
Austenitic Stainless Steel . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 41
Aluminum Alloys . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 46
General Considerations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 46
2000 Series . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 46
3000 Series . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 50
5000 Series . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 52
6000 Series . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 52
7000 Series . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 57
Weldability . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 61
Thermal Properties . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 63
Recommendations for LNG Piping . . . . . . . . . . . . . . . . . . . . . . . . 63
Nickel-Base Alloys . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 65
Nickel Steels . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 67
3-1/2 Percent Nickel Steel . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 67
9 Percent Nickel Steel . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 71
lnvar . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 78
Recommendations for LNG Pipeline . . . . . . . . . . . . . . . . . . . . . . . . . . . . 81
Quenched and Tempered Steels . . . . . . . . . . . . . . . . . . . . . . . . 81

INSULATION MATERIALS . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 85
Cellular Insulation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 85
Urethane Foams . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 85
Polystyrene Foams . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 86
Epoxy Foams . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 90
Silicone Foams . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 93
Phenolic Foams . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 94
TABLE OF CONTENTS
(Continued)

Page

Syntactic Foams . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 96
Cellular Glass . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 96
Cork and Balsa Wood . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 97
Vacuum Insulations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 98
High Vacuum . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 98
Evacuated Powder . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 99
Multilayer . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 101

VAPOR BARRIER . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 102

MISCELLANEOUS EQUIPMENT . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 106

Cryogenic Pumping Equipment . . . . . . . . . . . . . . . . . . . . . . . . . . . . 106


Valves . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 107
Instrumentation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 107
Temperature . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 107
Pressure . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 108
Flow . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 108

DISCUSSION . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 109

RECOMMENDATIONS FOR FUTURE WORK . . . . . . . . . . . . . . . . . . . . . . . 113

Additional Work on Design . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 113


Laboratory Determination of Mechanical and Thermal
Properties of Several Material Candidates . . . . . . . . . . . . . . . . . . . . . 114
Full-Scale Fracture Tests to Evaluate Resistance to
Fracture Initiation and Propagation . . . . . . . . . . . . . . . . . . . . . . . 114
Athens Loop . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 115

REFERENCES . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 116
APPENDIX

MATHEMATICAL DERIVATION FOR TEMPERATURE PROFILE ALONG AN


LNG PIPELINE . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . A-1
LIST OF FIGURES

Page

Figure 1. LNG Temperature Increase and Pressure Drop as a Function of Length for
Flow Through a lo-Inch-Diameter Pipe . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13
Figure 2. LNG Temperature Increase and Pressure Drop as a Function of Length for
Flow Through a lo-Inch-Diameter Pipe . . . . . . . . . . . . . . . . . . . . . . . . . . . 14

Figure 3. LNG Temperature Increase and Pressure Drop as a Function of Length for
Flow Through a 20-Inch-Diameter Pipe . . . . . . . . . . . . . . . . . . . . . . . . . . 15

Figure 4. LNG Temperature Increase and Pressure Drop as a Function of Length for
Flow Through a 20-Inch-Diameter Pipe ............................ 16

Figure 5. Pumping- and Cooling-Station Spacing for Case A . . . . . . . . . . . . . . . . . . . . 18

Figure 6. Pumping- and Cooling-Station Spacing for Case B . . . . . . . . . . . . . . . . . . . 18

Figure 7. Pumping- and Cooling-Station Spacing for Case C . . . . . . . . . . . . . . . . . . . . 19

Figure 8. Pumping- and Cooling-Station Spacing for Case D . . . . . . . . . . . . . . . . . . 19

Figure 9. Horsepower Requirements for Case A .......................... 23

Figure 10. Horsepower Requirements for Case B . . . . . . . . . . . . . . . . . . . . . . 24

Figure 11. Horsepower Requirements for Case C . . . . . . . . . . . . . . . . . . . . . . . 25


Figure 12. Horsepower Requirements for Case D . . . . . . . . . . . . . . . . . . . . . . . . 26

Figure 13. Plan View of a 6-Inch LOX Line . . . . . . . . . . . . . . . . . . . . . . . . . . . 29

Figure 14. lnvar Pipe With Mechanical Insulation . . . . . . . . . . . . . . . . . . . . . . . . 30

Figure 15. lnvar Vacuum-Jacketed Transfer Line . . . . . . . . . . . . . . . . . . . . . . . . . 31

Figure 16. Expansion Joint With Concrete Enclosure . . . . . . . . . . . . . . . . . . 35

Figure 17. Tensile Properties of Annealed Types 304 and 304L Stainless Steel . . . . 42

Figure 18. Tensile Properties of Cold-Worked Types 304 and 304L Stainless Steel . . . . 42

Figure 19. Reciprocating Beam Fatigue Strength of Cold-Worked Type 304 Stainless Steel . . 43

Figure 20. Impact Properties of Type 304 Stainless Steel . . . . . . . . . . . . . . . . . . 45

Figure 21. Coefficient of Linear Expansion of Several Materials . . . . . . . . . . . . . 45


Figure 22. Tensile Properties of 2000 Series Aluminum Alloys in the T6 Condition . . . . . 51

Figure 23. Effect of Heat Treatment on 2219 Aluminum Alloy . . . . . . . . . . . . . . . . 51


Figure 24. Effect of Welding on the Tensile Properties of 2014 and 2219 Aluminum Alloys
in the T6 Condition . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 53
Figure 25. Minimum Tensile Strengths of Some Annealed 5000-Series Aluminum Alloys . . . 53
LIST OF FIGURES
(Continued)

Page

Figure 26. Tensile Properties of 5000 Series Aluminum Alloys in the H3X Condition . . . 55

Figure 27. Effect of Welding on the Tensile Properties of 5052 Aluminum Alloy in the
H32 Condition .............................................. 55

Figure 28. Effect of Welding on the Tensile Properties of 5086 Aluminum Alloy in the
H34 Condition ................................................. 55

Figure 29. Effect of Welding on the Tensile Properties of 5456 Aluminum Alloy in the
H343 Condition . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 55

Figure 30. Charpy V-Notch Impact Energy of 5052-F Aluminum . . . . . . . . . . . . . . . 56

Figure 31. Charpy V-Notch Impact Energy of 5086-0 Aluminum . . . . . . . . . . . . . . . 56

Figure 32. Charpy V-Notch Impact Energy of 5154-0 Aluminum . . . . . . . . . . . . . . . 56

Figure 33. Tensile Properties of 6061 Aluminum Alloy . . . . . . . . . . . . . . . . . . . . . . . 58

Figure 34. Elongation of 6061 Aluminum Alloy . . . . . . . . . . . . . . . . . . . . . . 58

Figure 35. Charpy Keyhole Impact Energy of 6061-T6 Aluminum Alloy . . . . . . . . . 58

Figure 36. Tensile Properties of 7000 Series Aluminum Alloys in the T6 Condition . . . . 59

Figure 37. Effect of Welding on the Tensile Properties of 7002 Aluminum Alloy in the
T6 Condition . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 59

Figure 38. Tensile Properties of 0.75-Inch Aluminum Alloy 7039-T6 Plate, from 75 to -423 F . . 60

Figure 39. Coefficients of Expansion of Some Aluminum Alloys . . . . . . . . . . . . . . 64

Figure 40. Specific Heat of Some Aluminum Alloys . . . . . . . . . . . . . . . . . 64

Figure 41. Thermal Conductivities of Some Aluminum Alloys . . . . . . . . . . . . . . 64

Figure 42. Strength of Monel . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 66

Figure 43. Elongation of Monel . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 66

Figure 44. Strength of lnconel . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 66

Figure 45. Elongation of lnconel . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 66

Figure 46. Charpy V-Notch Impact Energy of Monel . . . . . . . . . . . . . . . . . . . . . . 68

Figure 47. Charpy V-Notch Impact Energy of Inconel ......................... 68

Figure 48. How Nickel Affects Keyhole-Notch Charpy Impact Test Properties ....... 69

Figure 49. Effect of Heat Treatment and Plate Thickness on Keyhole-Notch Impact
Properties of 3-1/2 Percent Nickel Steel Plate . . . . . . . . . . . . . . . . . . . . . . 69

Figure 50. Charpy Keyhole-Notch Impact Values for 3-1/2 Percent Nickel Steel Plate . . . 70
LIST OF FIGURES
(Continued)

Page

Figure 51. Charpy V-Notch Impact Values for 3-1/2 Percent Nickel Steel Plate ... 70
Figure 52. Relation of Tensile Properties to Temperature for Normalized and Stress-
Relieved 9 Percent Nickel-Steel Plate ............................ 73
Figure 53. Relation of Tensile Properties to Temperature for 9 Percent Nickel-Steel
Plate Welded With Inconel-Type Wire and Tested As Welded . . . . 73

Figure 54. Effect of Heat Treatment and Plate Thickness on Keyhole-Notch Charpy
Impact Properties of 9 Percent Nickel Steel . . . . . . . . . . . . . . . . . . . . . . . 74

Figure 55. Charpy Keyhole-Notch Impact Value for 9 Percent Nickel-Steel Plate . . . . 74

Figure 56. Results of 3/4-Width Charpy V-Notch Impact Tests on 3/8-Inch-Thick


Plates of 9 Percent Nickel Steel . . . . . . . . . . . . . . . . . . . . . . . . 76

Figure 57. S-N Curves for Notched and Smooth Rotating Beam Fatigue Specimens
of Normalized and Stress-Relieved 9 Percent Nickel-Steel, Room-
Temperature Tests . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 76

Figure 58. Typical Mechanical Properties of lnvar Plate . . . . . . . . . . . . . . . . . 79

Figure 59. Stress-Strain Diagram for Invar . . . . . . . . . . . . . . . . . . . . . . . . . 80

Figure 60. Modulus of Elasticity of lnvar ............................... 80

Figure 61. Stress-Strain Curves for HY-80 and 5 Ni-Cr-Mo-V Steel . . . . . . . . . . . 83


Figure 62. Charpy V-Notch Impact Test Results for HY-80 and 5 Ni-C-Mo-V Steels . . . 83
Figure 63. Variation of Mechanical Properties With Density of Urethane Foam ...... 87
Figure 64. Thermal Conductivity of Urethane as a Function of Density ......... 87

Figure 65. Effect of Temperature on Thermal Conductivity for Various Insulation


Materials .................................................... 88
Figure 66. Tensile Properties of Urethane Foams as a Function of Temperature . . . 88

Figure 67. Thermal Conductivity of Molded Polystyrene Foam as a Function of Density . . . 91


Figure 68. Thermal Conductivity of Polystyrene Foams . . . . . . . . . . . . . . . . . . . . 91
Figure 69. Relationship Between Compressive Strength and Density of Prefoamed
Epoxies . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 92

Figure 70. Tensile and Shear Strength of Epoxy Foam as a Function of Temperature . . . 92

Figure 71. Strength Properties of Phenolic Form as a Function of Density ........ 95


Figure 72. Relationship Between the Thermal Conductivity of Phenolic Foams and
Density .................................................... 95
LIST OF FIGURES
(Continued)

Page

Figure 73. Thermal Conductivity of Corkboard as a Function of Temperature ............. 97

Figure 74. Opacified Powder Performance . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 100

Figure 75. Performance of Evacuated Powder as a Function of Vacuum . . . . . . . . . . . 100

Figure 76. Performance of Multilayer Insulation as a Function of Vacuum . . . . . . . . . . . 101

Figure 77. Linear Expansion as a Function of Temperature for Three Vapor-Barrier


Materials . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 104

Figure 78. Effect of Temperature on Shear Strength of a Polyurethane Adhesive . . . . . . . 105

Figure 79. Pressure and Flow-Rate Capabilities of Commercially Available


Cryogenic Pumps . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 106
LIST OF TABLES

Page

Table 1. Heat-Transfer Resistances of Insulation and Ground . . . . . . . . . . . . . . 11

Table 2. Transfer of LNG Through a 10-Mile-Long Pipeline . . . . . . . . . . . . . . . . 17

Table 3. Expansion-Loap Spacing far Various Materials . . . . . . . . . . . . . . . . . . 33

Table 4. Allowable Design Stresses . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 36

Table 5. Pressure Design Thicknesses Required for a 10-Inch-Diameter Pipe Subjected


to a 1000 Psi Pressure . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 37

Table 6. Acoustic Velocities in Liquid Methane . . . . . . . . . . . . . . . . . . . . . . . . 40

Table 7. Low-Temperature Impact Properties of Annealed Type 304 Stainless Steel ... 44

Table 8. Low-Temperature Impact Properties of Type 304 Stainless Steel After Prolonged
Exposure . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 44

Table 9. Low-Temperature Charpy V-Notch Impact Properties of Inert-Gas-Shielded Metal-Arc


Weld in 3-1/2-Inch Type 304 Stainless Steel Plate .................. 44

Table 10. Selected Thermal Properties of Type 304 Stainless Steel .............. 46

Table 11. Typical Tensile Properties of Aluminum Alloys at Cryogenic Temperatures . . . . . 47

Table 12. Aluminum Alloys for Cryogenic Applications . . . . . . . . . . . . . . . . . . . 49

Table 13. Results of Tensile Tests of Some 5000-Series Aluminum Alloys at


Subzero Temperatures . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 54

Table 14. Tensile Properties of 7039 Aluminum Alloy Parent Metal at 75 and -320 F . . . 61

Table 15. Mechanical Properties of 7039 Aluminum Alloy Welds at 75 and -320 F, As Welded . . 62

Table 16. Notch-Toughness Properties of 3-1/2 Percent Nickel-Steel Welded Plate at -150 F . . 71

Table 17. Selected Thermal Properties of 3-1/2 Percent Nickel Steel . . . . . . . . . . . 71

Table 18. Charpy Impact Test Results on As-Welded 5/16-Inch-Thick 9 Percent Nickel-Steel
Plate . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 75

Table 19. Charpy Impact Test Results on As-Welded 1/2-Inch-Thick 9 Percent Nickel-Steel
Plate . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 77

Table 20. Selected Thermal Properties of 9 Percent Nickel Steel .............. 77

Table 21. Typical Ambient-Temperature Properties of Annealed lnvar . . . . . . . . . . . 78

Table 22. Effect of Titanium-Modified Filler Metal on the Mechanical Properties of


Gas-Tungsten-Arc Welded lnvar Alloy Plate 1/4-Inch Thick . . . . . . . . . . . 82
Table 23. Percentage Composition of lnvar Base Metal and Weld Filler Metal . . . . . . . 82

Table 24. Typical Mechanical Properties of HY-80 Steel and 5 Ni-Cr-Mo-V Steel . . . . . . 84
LIST OF TABLES
(Continued)

Page

Table 25. Summary of Urethane-Foam Properties . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 89

Table 26. Properties of Polystyrene Foams . . . . . . . . . . . . . . . . . . . . . . . . . . . . 90

Table 27. Properties of Epoxy Foams . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 93

Table 28. Typical Properties of Silicone-Resin Foams . . . . . . . . . . . . . . . . . . . . . . . 94

Table 29. Physical Properties of Syntactic Foams . . . . . . . . . . . . . . . . . . . . . . . . . . . . 96

Table 30. Cellular-Glass Insulation Thermal and Strength Values . . . . . . . . . . . . . . . 97

Table 31. Physical Properties of Balsa Wood . . . . . . . . . . . . . . . . . . . . . . . . 98

Table 32. Tensile and Elongation Properties of Vapor-Barrier Materials at -320 F . . . . . . . . 103

Table 33. Tensile Properties of Composite Laminates at Room and Cryogenic


Temperatures . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 103
Table 34. Tensile Properties of Thin Plastic Films . . . . . . . . . . . . . . . . . . . . . . . . . 104
MATERIALS OF CONSTRUCTION
FOR USE IN AN LNG PIPELINE

INTRODUCTION

Increased activity in the field of liquefied natural gas (LNG) has brought to the
minds of many gas-industry representatives the possibility of pipelining natural gas in
its liquid, rather than gaseous, form. The advantage of liquefying the gas for transpor-
tation is that more standard cubic feet of natural gas can be carried through a given
cross-sectional area of pipe in comparison with pressurized transmission systems at
ambient temperature. Any advantage of increased throughput must, however, be
weighed against the necessary costs to initially liquefy the gas and to keep it cold through
insulation and reliquefaction (LNG pipelines will, most likely, be operated over a tem-
perature range of -259 to -140 F). Also, increased costs may be expected in order to
circumvent design problems resulting from the cryogenic temperature.

Aside from considerations of the possibility of competitive transfer by liquefying


the natural gas before transmission, projects are envisioned which may transfer already
liquefied LNG between storage depots over considerable distances. In general, three
broad areas in which the potential for transmitting LNG would appear to exist are

(1) Short imperative transfers, such as in-plant piping and lines from
storage to loading or unloading facilities

(2) Optional transport methods, where a choice between trucking,


barging, or pipelining LNG might exist (e. g., transporting LNG
from a central liquefaction plant to satellite storage depots or
between storage depots)

(3) Base-load operations where transmission of LNG might be con-


sidered to markets over relatively long distances.

The objective of the PRCI Pipeline Research Committee’s Project PR-3-42,


being reported here, is to survey commercially available materials which are con-
sidered acceptable or potentially acceptable from a technical viewpoint for the construc-
tion of an LNG pipeline. Although material-cost figures are given and some discussion
of them is presented, the scope of the current work does not include an evaluation of the
economic feasibility of pipelining LNG. Because of the many evident alternatives in the
actual design and construction of an LNG pipeline in addition to the many variables and
uncertainties involved in economic studies in various parts of the country and the world,
it was agreed that this investigation would summarize the data and information on mate-
rials of construction so that such economic studies could be made.

For “ground rules” in the material survey, flow rates from 200 to 500 MMCFD are
assumed. By crude analysis , pipe sizes from 10 to 20 inches in diameter were deemed
appropriate for these flow rates. There are several possibilities for operating conditions
due to the fact that the critical pressure and temperature for LNG (actually liquid
2

methane) are 673 psia and -116 F. * The choice of operating conditions will depend on
the nature of the particular application. For purposes of this report, initial pressure
and temperature conditions at the beginning of an LNG pipeline were assumed to be
atmospheric pressure and -259 F. Also, in this report, it is assumed that the LNG
pipeline is level.

In the survey of construction materials, major attention has been given to pipe
materials. Materials which have transition temperatures above the expected service
temperature have not been ruled out because of the special relation between estimated
fracture speed and speed of decompression in LNG. This has allowed consideration of
several of the less expensive materials. Further justification for considering mate-
rials with transition temperatures above -259 F lies in the fact that LNG pipelines will
have a temperature profile starting at -259 F and ending at a temperature somewhat
lower than -120 F. Hence, it is possible that fracture-toughness properties can be
“telescoped” commensurate with the temperature profile of the LNG pipeline.

Important mechanical properties of the pipe materials are presented where


available. Insulating materials and techniques are also reviewed and their properties
given. Vapor-barrier materials and, to a lesser extent, miscellaneous equipment are
also treated.

Because materials of construction are somewhat dependent on transmission dis-


tance, particularly insulations, studies aimed at determining a “modular distance” over
which LNG can be transmitted without additional horsepower input, i. e., pumping or
reliquefaction, have been made. These studies are applicable to the second and third
transmission systerms mentioned above. Long-distance transmission, of course, con-
sists of a string of “modules”. In this report, transmission distances, based on
steady-state analyses, are given as a function of volume flow rate, heat-influx, pres-
sure drop, and pipe size.

Materials of construction are often tied in with the design approach. For this
reason, some basic design configurations are described which illustrate the combined
use of some candidate materials.

Finally, in carrying out the work on this project during 1966, some future steps
in the research and development of an LNG pipeline became evident. These are identi-
fied in the last section of the report.

In the case of methane, there is no pressure at temperatures higher than -116 F which will cause the gas to stay in the liquid
state.. Thus for purposes of this study, -116 F is the highest temperature of concern.
3

SUMMARY

A survey of current literature has been carried out to determine the technical
feasibility of transmitting LNG over considerable distances. A steady-state heat-
transfer analysis was developed which gives an indication of the spacing between cool-
ing and pumping stations for various pipe diameters, insulations, and flow rates.

Modular Distance Studies

The results of the heat-transfer analysis indicate that LNG can be transmitted
over fairly long distances without intermediate pumping and cooling stations. F o r i l -
lustrative purposes, two examples are considered with the following conditions:

The calculations indicate that the pumping- and cooling-station spacing for Example 1
would be 25 and 100 miles, respectively, whereas, for the conditions of Example 2, it
appears that pumping and cooling stations would be required every 130 and 175 miles,
respectively. The pumping-station spacing is based on a pressure drop of 300 psi,
whereas a cooling station is deemed necessary wherever the LNG temperature reaches
-140 F.

Horsepower requirements for pumping and cooling are based on the assumption
that LNG is initially available at -259 F and 1 atmosphere pressure. The mechanical
power required to raise the LNG pressure from 0 to 1000 psig at the initial pump station
is calculated to be approximately 1300 hp for Example 1 and 3250 hp for Example 2
(pump efficiency = 0.75). For a flow of 200 MMCFD through a lo-inch line, it takes
15.5 hp to overcome the pressure drop per mile of pipeline, whereas for a flow of
500 MMCFD through a 20-inch-diameter pipe it takes only 7.4 hp/mile. Assuming a
10 percent cooling station efficiency, the calculations indicate that the cooling horse-
power requirements are 1137 hp/mile (Example 1) and 1875 hp/mile (Example 2).

It is evident from the above figures that the largest power requirements are for
cooling. Consequently, it may be economically advantageous to either increase the
distance between cooling stations by increasing insulation thickness, or to keep the
line length below that which would require an intermediate cooling station.
4

Design Considerations

Piping systems for transporting liquefied natural gas are not basically different
from those for the more familiar fluids, However, due to the temperature range en-
countered - ambient to -259 F, it is necessary to provide for the longitudinal thermal
contraction of the line. This can be best accomplished by using either Invar (36 percent
nickel steel) , expansion loops, or expansion joints (bellows).

The LNG line can be constructed without expansion joints by using materials,
such as Invar, which have a low coefficient of expansion. The longitudinal stress of
an Invar line at operating temperature for a fully restrained line is approximately
6400 psi. The pipe layout and anchor locations for Invar pipe would be determined in
the same manner as for conventional gas transmission lines.

Expansion-loop design for an underground line requires the construction of a


chamber to allow free movement of the loop piping. The underground chamber design
would be similar to the designs presently used for steam or chilled water lines, al-
though several modifications would be necessary. For a lo-inch line and a specified
loop height of 20 feet and loop width of 10 feet, the number of expansion loops required
per mile for various materials is as follows: 6061-T6 aluminum (31.3), 5083-0 alu-
minum (19. 6), Type 304 stainless steel (21.2), 3-1/2 percent nickel steel (15.9), and
9 percent nickel steel (8. 7). Because the loop spacing is limited by the longitudinal
stresses in the main line outside the loops, allowable stresses for thermal contraction
were determined in accordance with USASI B31.3.

In-line expansion joints of stainless steel bellow construction are available for
cryogenic service. Expansion joints are not recommended for an LNG line operating
at 1000 psi because bellows subjected to this pressure become expensive and have
limited axial movement. For line pressures less than 250 psi the amount of axial
movement and the cost of bellows becomes more reasonable. A line constructed from
9 percent nickel steel would require 37 bellows per mile, where the design axial move-
ment of each bellow is computed as 2.94 inches. The number of bellows required per
mile of pipeline increases to 53 for Type 304 stainless steel and to 69 for aluminum,

The code for gas-transmission lines, USASI B31.8, is not specifically intended to
cover the pipe materials and operating temperatures of LNG pipelines, Although al-
lowable design stresses for thermal contraction of the line were based on USASI B31.3
and the pipe-wall-thickness calculations were based on USASI B31.8, it is not intended,
in this report, to infer an allowable design stress for LNG pipelines; this is a matter
requiring further considerations,

Pipe Materials

The literature survey indicated that there are a large number of materials which
are either acceptable or potentially acceptable for the construction of an LNG line.
While it has been recognized that economic factors are important in the construction of
an LNG line, a number of materials have been included in the study that may not be ap-
plicable because of high cost.
5

Fracture-toughness properties are important from two viewpoints: propagation


resistance and initiation resistance. If a fracture should start, propagation resistance
should keep it as short as possible. Because the decompression velocity of LNG is
expected to be high (perhaps higher than any possible fracture speed), all fractures
might be expected to be short, whether brittle or ductile. If this is true, it allows con-
sideration of materials which have transition temperatures below the service tempera-
ture. Furthermore, because the LNG pipeline will have a pronounced temperature pro-
file, it may be possible to “telescope” fracture-toughness properties, using higher
transition temperature materials at the warmer temperatures. Therefore, the over-
riding criteria for material acceptance from a fracture-toughness viewpoint might be
its initiation resistance.

Austenitic stainless steels meet many of the requirements for low-temperature


service. Of these stainless steels, Type 304 is probably the most applicable. Available
data have indicated adequate strength and fracture toughness. The greatest limitation
is cost. Basic-material costs may be expected to be on the order of eight to nine times
the cost of ordinary carbon steels, e. g., A53 steel.

Aluminum alloys are considered good candidates for use in an LNG pipeline. I t
appears that good strength can be achieved and resistance to fracture propagation ap-
pears satisfactory, at least for several alloys. Fracture-initiation resistance, how-
ever, is uncertain and will have to be determined. Of the alloys examined in the
report, 5456-0, 5083-0, 6061-T6 and 7002-T6 appear to be the most promising.

The possible use of aluminum is attractive from an economical viewpoint consider-


ing base costs of material only. However, additional costs are imposed due to design
considerations such as the requirement of flanged connections if expansion joints (stain-
less steel) are incorporated in the line. Furthermore, because aluminum has a large
thermal coefficient of expansion, additional compensation in the form of expansion loops
or expansion joints are necessary over that required for other materials such as nickel
steels. Aluminum base-material costs may be expected to be on the order of three to
four times the base costs of an ordinary carbon steel; however, these costs do not
acknowledge any difference in allowable stress levels,

Both Inconel and Monel have excellent properties at low temperatures and would
be candidates for an LNG pipeline if they were not so expensive. Unless overall costs
are significantly reduced by the fact that expansion bellows or loops are not needed,
their use does not appear to be justified from economic considerations.

Nickel has been shown to be outstanding for lowering the transition temperature of
ferritic steels. Of the three nickel steels (3-1/2 percent, 9 percent, and Invar) con-
sidered in this report, possibly the most applicable for LNG pipelines would be the
9 percent nickel steel in the quenched and tempered condition. This is primarily based
on strength, toughness, and cost considerations. Invar is perhaps better from the
fracture-toughness viewpoint, but the 9 percent nickel steel is satisfactory. Invar,
aside from its higher base-material costs (ten to twelve times that for carbon steel),
would have lower installation costs because expansion loops or bellows would not be re-
quired. The 3-1/2 percent nickel steel, while it does not have the fracture-toughness
levels that 9 percent nickel does, is significantly less expensive. Base-material costs
for 3-1/2 percent and 9 percent nickel steels are about three to three and one half times
and six to seven times those for ordinary carbon steel, respectively. The 3-1/2 percent
6

nickel steel may be qualified for LNG pipeline use by additional fracture-toughness
testing.

Some of the present quenched and tempered steels such as 5Ni-Cr-Mo-V and
HY-80 may be suitable for LNG pipeline service if it can be demonstrated that

(1) their fracture speeds are sufficiently below the decompression


velocity of LNG

(2) their resistance to fracture initiation is sufficient.

Insulation Materials

As well as pipe-construction materials, suitable low-temperature insulation mate-


rials were also investigated. The insulations covered in this report fall into two
classes:

(1) nonvacuum or mechanical-type insulation

(2) vacuum insulation.

Of the conventional insulations investigated, urethane-foam insulation appears to have


the best combination of thermal-performance characteristics, strength, ease of fabri-
cation, and low cost. For an equivalent thermal resistance to heat transfer as provided
by 2 inches of urethane foam on a 10-inch-diameter pipe, it would require 2.40 inches
of phenolic foam, 2.83 inches of polystyrene foam, 3.25 inches of cork, 3.75 inches of
epoxy foam, 6.25 inches of balsa wood, and 6.25 inches of cellular glass. Because
urethane foam is available either in boards or billets as well as foam-in-place formula-
tions, the insulation can be either factory or field installed.

Vacuum-type insulations, which are, in effect, two concentric shells with a


vacuum in the annulus, have been used for many years to insulate tanks for the storage
of liquefied oxygen, nitrogen, and hydrogen. In such insulations the heat transport due
to gas conduction and convection is reduced (to a negligible amount in most cases) by
removing most of the gas from the insulation space,

Although vacuum-jacketed insulations have thermal conductivities which may be


more than 100 times lower than the conductivities of mechanical-type insulations, the
high cost of vacuum-jacketed insulations which may run as high as $200 per linear foot
for a lo-inch-diameter pipe may make this insulation uneconomical for the construction
of an LNG line. Due to the high installation cost and probable requirement for periodic
pumping to maintain the high vacuum, vacuum-jacketed insulations are not recommended
for an underground LNG pipeline.

Vapor Barrier

The primary problem associated with the use of conventional insulations stems
from the requirement of a vapor barrier. Metals, foils, plastics, and various types of
7

coatings may be used, although the most practical and most widely used vapor barriers
are a combination of mastics and membranes. One installation used hot asphalt and a
10 by 10 mesh-impregnated glass cloth applied directly over a cellular-glass insulation.

Miscellaneous Equipment

Cryogenic pumps, valves, and instrumentation are either presently available or


can be readily designed and built using present state-of-the-art technical knowledge and
capabilities.
8

MODULAR DISTANCE STUDIES

Transmission of cryogenic fluids has been limited up to the present time to above-
ground vacuum-jacketed lines. Lines up to 12 inches in diameter and 1-1/2 miles in
length have been built. No references have been found which indicate the construction of
a cryogenic transfer line in an underground location.

The feasibility of transporting liquefied gases over appreciable distances through


piping systems has been discussed by Jacobs(l). The heat-transfer calculations which
are given in this section of the report are more applicable to the long-distance transfer
of LNG because the heat leak is assumed to be a function of distance along the line,
whereas it was assumed to be constant in the calculations carried out by Jacobs. In the
analysis that follows, a steady-state condition and single-phase transfer through the
line are assumed.

For transfer rates of 200 and 500 MMCFD through lo-inch and 20-inch-diameter
pipes, respectively, characteristics such as the temperature profile and pressure drop
were determined as a function of distance for various thicknesses and types of insula-
tion. The curves thus generated can be used for predicting the terminal LNG tempera-
ture and the pumping- and cooling-station spacing.

Mathematical Analysis

The derivation of a formula for predicting the temperature profile along an LNG
pipeline is presented in the appendix. From the appendix, the equation relating the
LNG temperature to the pipeline length is as follows:

(A- 17)

where L= pipeline length between points a and b, ft

W = LNG weight flow rate, lb/sec

c = specific heat, Btu/(lb)(F)

R 2 3 ’ = heat-transfer resistance of the insulation, (sec)(ft)(F)/Btu

R34’ = heat-transfer resistance of the ground, (sec)(ft)(F)/Btu

T 4 = ground-surf ace temperature, F

f = friction factor
9

V = fluid velocity, ft/sec

r1 = pipe inside radius, ft

Tb-Ta = temperature rise over length L from point a to point b, F

g = gravitational acceleration, ft/sec2

J = 778 ft-lb/Btu.

The mechanical power required to pump LNG through a pipeline of length L is


also given in the appendix. It is

(A- 18)

where p a -p b = pressure drop over length L, psfa

w = specific weight of LNG, lb/ft 3

e = pump efficiency.

Calculations

Heat-Transfer Resistance

Prior to using Equation (A-17) for calculating pumping- and cooling-station spac-
ing, the total heat-transfer resistance of insulation and ground (R23’ t R34’) were deter-
mined for three types of insulation which encompass a broad range of available thermal
conductivities, The three types of insulations considered in this section were urethane
foam, evacuated perlite ( 10-2 mm mercury) and evacuated multilayer insulation
(10-4 mm mercury). Although a more complete discussion of insulation materials is
presented in a later section of this report, the thermal conductivities tabulated below
are typical values for temperature boundary conditions of -260 F and +70 F for the
densities indicated:

The pressure for high-vacuum insulation should be 10 -4 mm mercury or less, while a


pressure of 10 - 2 mm mercury is satisfactory for evacuated powder.
10

The soil thermal conductivity (k34 ) was taken as 0.62 Btu/(hr)(ft)(F). This value
represents frozen soil containing normal moisture contents. For dry sand and dry soil
including stones, the thermal conductivities are approximately 0.19 and 0.30 Btu/(hr)
(ft)(F).

Total heat-transfer resistance for lo-inch and 20-inch-diameter pipe insulated


with 2, 4, and 6 inches of urethane foam were calculated for the above insulation and
soil thermal conductivities. Similar calculations were made for 2 and 4 inches of
perlite and 2 inches of multilayer insulation. The centerline of the pipe was taken as
3 feet below ground surface. Heat-transfer resistances of insulation and ground are
expressed in the appendix as

(A-7 and A-8)

The results of the calculations of R23’ and R34’ as listed in Table 1 indicate that

(1) The contribution of the total heat-transfer resistance of the soil


is a significant percentage when urethane is used for insulating
the pipe, but is insignificant for the other two insulating materials.

(2) For each additional 2-inch increase in urethane-insulation thick-


ness, the corresponding change in the heat-transfer resistance
decreases. In going from 2 to 4 inches of urethane insulation on
a lo-inch-diameter pipe, R 23 ’ increasedby 1.20 x 10 4 (sec)(ft)(F)/Btu,
whereas in going from 4 to 6 inches of insulation, the increase in
R 23 ’was 0.96 x 10 4 (sec)(ft)(F)/Btu.

(3) If the total heat-transfer resistance of 2 inches of urethane insula-


tion on a lo-inch-diameter pipe is taken as base, then, for similar
conditions, evacuated perlite and multilayer insulation will have
total heat-transfer resistances approximately 12 and 422 times
greater than urethane insulation, respectively.

Pumping- and Cooling-Station Spacing

Equation (A-17) was used to calculate pipeline length as a function of LNG tem-
perature and pressure drop for two pipe sizes and for various insulations. Calculations
were based on flows of 200 MMCFD (through a lo-inch-diameter pipe) and of
500 MMCFD (through a 20-inch-diameter pipe) of natural gas, These combinations may
not represent the optimum conditions of flow and pipe-diameter from either an econom-
ical or a flow standpoint. However, these combinations were based on crude estimates
and should be at least a first approximation.

The specific weight of LNG, the specific heat of LNG, and the friction factor
must be known before Equation (A- 17) can be used to calculate L. The specific weight,
w, was taken to be 26.3 lb/ft3 and was assumed to be constant. An average value, over
the temperature range of interest, of the specific heat of LNG, c, was taken as 0.9
Btu/(lb)(F). The friction factor, f was estimated to be approximately 0.014.
11

TABLE 1. HEAT -TRANSFER RESISTANCES OF INSULATION AND GROUND


12

Figures 1 through 4 present the results obtained from Equation (A-17) for the
insulations and pipe diameters listed in Table 1. The pressure drops given in these
figures were calculated from Equation (A-14). For the insulations considered, it is
evident from Figures 1 through 4 that the pressure-drop curve will be the governing
curve for the majority of the applications. This means that more pumping stations will
be required than cooling stations.

Figures 1 through 4 will be used to discuss the pumping and cooling requirements
for three types of transfer of LNG. The three types considered are

(1) Short imperative transfers (less than 1 mile), such as in-plant


piping and lines from storage to loading or unloading facilities

(2) Optional transport methods (1 to 10 miles) where a choice be-


tween trucking, barging, or pipelining LNG might exist

(3) Base-load operations where long-distance LNG transmission


might be considered.

For the following discussion it is assumed that initially at the suction end of the
pump LNG is available at approximately -259 F and 1 atmosphere. Furthermore, for
each of the three transmission distances mentioned above, only the following four cases
will be examined in detail because the other type and other thicknesses of insulation as
given in Table 1 will exhibit behavior somewhere between the two extremes for each
pipe diameter listed below:

For short imperative transfers of 1 mile or less, Figures 1 to 4 indicate that the
temperature rise due to heat influx and fluid friction will be less than 5 F. Thus, the
LNG terminal temperature will be approximately -254 F. In order to prevent flashing
at the terminal end of the line for LNG at -254 F, the absolute pressure should be
around 20 psia. Because Figure 1 indicates that the pressure drop will be less than
30 psi for a 1-mile transfer line, the pump discharge pressure should be from 30 to
40 psig. For a short transfer line of this distance, there is little advantage in using
the high-vacuum multilayer insulation, and 2 inches of urethane insulation on either a
lo-inch or a 20-inch-diameter line would be sufficient.

Similar calculations for a 10-mile-long transfer line indicate in Table 2 that very
little would be gained from using the best insulations currently available because the
pressure drop and not the temperature increase is the governing criterion.
17

TABLE 2. TRANSFER OF LNG THROUGH A 10-MILE-LONG PIPELINE

For the third transmission distance under consideration - a long transmission line
designed to carry LNG over several hundredmiles - the maximum line pressure and the
allowable pressure drop between pumping stations must be specified. The maximum
pump pressure and the pressure drop were taken as 1000 psi and 300 psi, respectively.
This means that wherever the line pressure drops to 700 psi, a pumping station will be
required. Because the line pressure is always maintained at or above 700 psi, there
is little danger of LNG flashing as long as the liquid temperature is below -116 F.
For this temperature the saturation pressure is about 673 psia.

As is pointed out in the appendix the temperature increase will be less than 10 F
for a pump pressure difference of 1000 psi and a pump efficiency of 50 percent, Several
manufacturers have indicated that for the pump sizes and capacities considered, pump
efficiencies in the vicinity of 75 percent can be realized. That efficiencies of this order
can be achieved for LNG pumps is uncertain and needs to be confirmed by test,

In view of this discussion it appears reasonable to assume that the temperature


increase at the first pumping station will be less than 10 F for a pressure increase of
1000 psi. At the additional pumping stations required to overcome the pressure drop of
300 psi, the temperature increase due to raising the pressure should be less than 5 F.

Therefore, at the discharge side of the first pumping station the LNG is at a
pressure of 1000 psi and at a temperature of about -249 F. The pressure-drop curve
in Figure 1 indicates that for a pressure drop of 300 psi, the distance between pumping
stations should be approximately 25 miles. For Case A (2 inches of urethane on a 10-
inch-diameter pipe) the temperature increase for 25 miles is from -249 F to -222 F.
At the pump discharge of the second station the LNG is now at 1000 psi and -217 F.
Because the pressure drop curve is linear, the pump station spacing would be again
25 miles. However, during this interval the LNG temperature has increased from
-217 F to -192 F. A continuation of this procedure would indicate the spacing between
cooling stations. The procedure can be followed to determine at what point along the
line LNG would require additional cooling if flashing of liquid is to be prevented. In
order to provide a margin of safety, it was arbitrarily decided that a cooling station
would be needed whenever the LNG temperature approached -140 F, although the satura-
tion temperature at 700 psi is about -116 F.

Figures 5 through 8 present the pumping and cooling station spacing for Cases A
through D. Figure 5 indicates that for a flow of 200 MMCFD through a lo-inch-diameter
18

FIGURE 5. PUMPING- AND COOLING-STATION SPACING FOR CASE A

FIGURE 6. PUMPING- AND COOLING-STATION SPACING FOR CASE B


19

FIGURE 7. PUMPING- AND COOLING-STATION SPACING FOR CASE C

FIGURE 8. PUMPING- AND COOLING-STATION SPACING FOR CASE D


20

pipe insulated with 2 inches of urethane, a pumping station would be required every
25 miles, and a cooling station would be required every 100 miles. For Case B
(Figure 6) the pumping-station spacing is the same as for Case A, because the flow and
pipe diameter are identical. However, the distance between cooling stations has in-
creased to 375 miles. For transmission lines under 100 miles it appears that Case A
is the more economical system, because intermediate cooling stations are not required
for either Case A or Case B, but the construction and material costs would be much
greater for a high-vacuum-jacketed line than for a line mechanically insulated.

Figure 7 indicates that for Case C the pumping and cooling station spacing is 130
and 175 miles, respectively. LNG enters the first cooling station at -140 F and
895 psi, and is discharged at -259 F and 895 psi. The pressure drop from 895 to
700 psi locates the third pumping station 83 miles from the first cooling station. The
distance between cooling stations can be increased by increasing the thickness of
urethane insulation from 2 to 4 inches.

Because of the high-vacuum multilayer insulation used for Case D (Figure 8), a
cooling station is required every 1950 miles. For this case, the major heat sources
are the fluid friction along the pipe and the heat inputs at the various pumping stations.
A pumping station is required every 130 miles, the same as for Case C.

Horsepower Requirements

Mechanical power must be supplied to the pumps to overcome the pressure drop
along the pipeline. Also, additional power must be supplied to the cooling stations to
lower the LNG temperature from -140 F to -259 F.

The necessary mechanical power to operate the pumps can be calculated from
Equation (A- 18)

(A- 18)

Taking the pump efficiency, e, as 0.75, the horsepower requirements for Case A and a
1-mile test-section length for a pump-discharge pressure of 30 psig would be

Because both the l-mile and the lo-mile-long transfer lines do not require Intermittent
cooling stations, the power for cooling is taken as zero.

For Case A and a lo-mile test section, the recommended discharge pressure
from Table 2 is 135 psig. Thus, the power requirements are

Similar calculations were performed for a lo-mile transfer line for Cases B, C, and
D, and the results are tabulated below:
21

For long-distance transmission of LNG, the power requirements were broken


down into three categories:

(1) Pump power necessary to raise the pressure of LNG from 0 to 1000 psig
at the initial pumping station

(2) Pump power required to overcome the 300 psig pressure drop between
stations

(3) Cooling power necessary to prevent LNG from flashing because of the
heat influx and the heat generated due to internal friction

The mechanical power required to raise the LNG pressure from 0 to 1000 psig
at the initial pump station for a flow of 200 MMCFD through a lo-inch line was
calculated to be

For a pressure drop of 300 psig, the distance between pumping stations for cases
A and B was 25 miles. Therefore, the mechanical power required to overcome the
pressure drop per mile of pipeline is

The total heat absorbed by the LNG due to the heat influx through the pipewalls
and internal friction is given by

This comes from the denominator of Equation (A-17).

The initial and terminal LNG temperature between cooling stations were assumed
to be -259 F and -140 F, respectively. Therefore, for 200 MMCFD flow through a 10-
inch-diameter pipe with 2 inches of urethane insulation (Case A), the heat influx is
calculated to be
22

Converting from Btu/(sec)(ft) to hp/mile, we get

For an assumed 10 percent cooling-station efficiency, the mechanical power required


for cooling becomes

Therefore, the total mechanical power for pumping and cooling is

Also, to this average figure must be added the power required at the initial pumping
station to raise the LNG pressure from 0 to 1000 psig. The results of these calcula-
tions for Case A are summarized in Figure 9. Similar calculations were carried out
for Cases B through D, and the results are summarized in Figures 10 through 12.

From Figure 9 it is evident that the largest power requirements for Case A are
for cooling. The pumping power represents less than 2 percent of the total power re-
quirements (cooling and pumping). Depending on economic considerations, the horse-
power requirements for cooling may be reduced if the thickness of urethane insulation
is increased. Therefore, in an actual pipeline design, the cost of increasing the
urethane insulation thickness must be weighed against the savings resulting from the
requirements for fewer cooling stations and reduced horsepower for cooling.

In Figure 10, the horsepower for pumping is 13.2 percent of the total power
required for cooling and pumping, Because the insulation used is the best presently
available, increasing the vacuum or the annular space will decrease the cooling require-
ments only slightly, since the major source of heat is internal friction generated by
LNG flowing along the pipe. The horsepower requirements for pumping may be reduced
by either decreasing the flow of LNG or by increasing the pipe size. However, the
economics of doing this is certainly questionable,
26

FIGURE 12. HORSEPOWER REQUIREMENTS FOR CASE D


27

DESIGN CONSIDERATIONS

Piping systems for transporting liquefied natural gas are not basically different
from those for the more familiar fluids. But, some differences must be considered
by the designer because of the temperature range encountered - ambient to -259 F.
These differences are due to differences in the properties of materials at low temper-
atures, the emphasis placed on the selection of an economic insulating system, and
the necessity of providing flexibility in the system due to thermal contraction of the
pipeline.

Insulations required for piping systems for liquefied gases range from no insula-
tion to the best vacuum-type insulation. The type of insulation used for a particular
application is determined by several considerations: allowable heat leak, material and
installation costs, operating and maintenance costs, ruggedness, length, and frequency
of use, Although the heat-transfer characteristics of various insulations will be
covered more fully in a later section, the two types of insulation possible will be briefly
introduced because they enter into the design considerations.

The insulations covered in this report fall into two classes: (1) nonvacuum or
mechanical- type insulation and (2) vacuum insulation. With the former, the insulation
is fabricated by shaping and fitting pieces of insulation to cover the pipe; the seams
between the pieces can be sealed by a suitable mastic, epoxy, etc., or by an external
covering. The foamed-in-place plastics which also belong to the first class of insula-
tions may not require a vapor barrier if they themselves have closed cells, do not have
fissures, and bond to the surfaces that they insulate. However, this cannot be relied
upon and vapor barriers of some sort appear to be advisable for the foamed plastics.
Vacuum insulations are in effect two concentric pipes with a vacuum in the annulus.
In such insulations the heat transport due to gas conduction and convection is reduced
(to a negligible amount in most cases) by removing most of the gas from the insulation
space, i. e. , by evacuation. Long-term permeation through the pipe or jacket and
outgassing (i e., release of gas from foreign materials within the vacuum space or
from the pipe or jacket metal) would probably require periodic vacuum pumping to
maintain heat losses below those for mechanical insulation.

An important factor - once the pipe size and system layout have been determined -
is providing for thermal contraction or flexibility of the system. There are four basic
approaches to the problem:

(1) Lay out the system so that it is inherently flexible

(2) Use of Invar (36 percent nickel steel)

(3) Use of expansion loops

(4) Use of expansion joints (bellows).

Methods (1), (3), and (4) are commonly used in ordinary piping systems, but be-
cause of vacuum-jacketed (V-J) pipe require additional considerations. The major ad-
vantage of Invar over other cryogenic pipe materials is its low coefficient of expansion-
0.9 x 10 -6 average from room temperature to -320 F. This amount of expansion is
low enough to allow the use of Invar pipe without expansion loops or joints.
28

An above-ground V-J pipe system which is installed in hilly, rugged terrain and
is more than 8400 feet long is illustrated in Figure 13, taken from an article by
Haettingert (2) . Because of the terrain it was necessary that the line have many turns.
Hence, these turns were used to provide system flexibility while using stainless steel
for the inner line. As can be seen, the approach used was to lay out the line as a
series of “L’s”, with the pipe being anchored at each end of the “L”. By sizing the
two legs so each “L” was not overstressed, flexibility was provided for the entire sys-
tem. Movement of the line during cooldown is as much as 20 inches in some places.
Hence, an elaborate support system of rollers and trolleys was required. Laying out
the system so that it is inherently flexible would be applicable only for certain hilly
terrain and would be difficult to accomplish for an underground piping system. Thus,
the last three approaches (i.e., use of Invar, use of expansion loops, and use of ex-
pansion joints or bellows) in providing piping flexibility due to thermal contraction will
be investigated more fully in the following discussion.

Fully Restrained line

The use of Invar pipe without expansion joints shows several practical advantages
over the other materials. This pipe has very low thermal stresses from expansion and
cooldown. For this reason Invar pipe construction is similar to conventional
transmission-line practice - no expansion joints or intermediate anchors are required
and the pipe can follow the earth surface contours to a reasonable degree. Invar has
been used satisfactorily in cryogenic applications although further work needs to be
done to determine realistic design stresses for this material.

The longitudinal thermal stress at an operating temperature of -259 F for a fully


restrained line calculated from

is 6030 psi. Young’s modulus, E, was taken as 21.0 x 10 psi, ∆ t was set equal to
6
-6
319 F (60 F to -259 F) and 0.9 x 10 was used for the mean coefficient of expansion, a.

Since Invar pipe does not move axially, factory foamed-in-place insulation can
be used on this pipe. This type of insulation has the advantage of eliminating the
numerous insulation joints when short half sections of insulation are field installed.
Also, field-construction time would be reduced by using preinsulated 40-foot lengths of
pipe.

Figure 14 shows a proposed design of Invar pipe with foamed-in-place insulation.


For a 40-foot length of pipe, 36 feet of the pipe are insulated with factory installed
foamed-in-place insulation complete with either a mastic or a light-gauge-metal vapor
b a r r i e r . After the girth weld is made in the field, foamed-in-place insulation and a
vapor barrier is applied to the ends of the pipe which were not factory insulated to
complete the installation.

A schematic diagram of an Invar V-J line is presented in Figure 15. This line is
concentrically placed in the jacket and supported by a low heat-conducting suspension
system such as Teflon, glass, ceramic, stainless steel pins or balls, or impregnated
32

fiber glass, thus reducing solid conduction. The openings between the inner line and
jacket at the ends of the jacket are sealed by a thin stainless steel seal having a long
heat path. Various methods are used to create a long-heat-path seal at these ends,
thus reducing heat loss by conduction.

Because girth welds must be made in the field, the factory-installed jackets and
seals terminate a few feet from each end of a 400-foot length of pipe. The annular space
is evacuated through the pump-out valves to an absolute pressure of less than 1 x 10-4
Torr (mm of Hg).

The integrity of the shop-fabricated-and-sealed vacuum-jacketed spool is generally


checked by

(1) Hydrostatically testing the inner line to 1-1/2 times the design
pressure

(2) Completely radiographing and inspecting all butt welds in the inner
line

(3) Leak testing with a mass-spectrometer-type leak detector before


shipment.

After the girth weld of the inner line is completed and radiographed in the field,
a short length of stainless steel pipe is welded to the factory-installed outer jacket.
The annular space in the vicinity of the girth weld is then evacuated to complete the
installation.

Expansion loops

In conventional piping systems, sufficient flexibility is obtained by the use of


expansion loops. The expansion-loop spacing required for uniform-thermal-cooldown
piping flexibility will be examined next.

Expansion-loop design for an underground line requires the construction of a


chamber to allow free movement of the loop piping. Underground-chamber construction
for steam or chilled-water lines is described in Reference 3. This design with several
modifications could be used in conjunction with a mechanical type of insulation such as
urethane or Foamglas.

Either the loop height, H, or the loop spacing, U’ (see sketch below) can be
determined by the approximate chart solution contained in Reference 4. The loop height
was specified as 20 feet, and the loop spacing U’ was calculated.

Allowable stresses for thermal contraction were determined in accordance with


USASI B31.3 as described under “Pipe Stresses” section of this report. The allowable
loop-contraction stress is 9000 psi for aluminum 6061-T6, 28, 125 psi for stainless
steel Type 304, and 32,475 for 3-1/2 percent nickel. No allowable design stresses for
pipe are given in USASI B31.3 for 5083-O aluminum and 9 percent nickel; the corres-
ponding values for these materials have been assumed to be 14,625 and 45,000 psi,
respectively, based on the ratio of allowable loop-contraction stress to allowable plate
stress for the materials listed above.
33

The loop layout used for the LNG line is shown below:

A constant loop configuration has been assumed for comparative purposes. Actually,
optimum configurations could be developed for each particular application.

The results of the calculations for various materials are listed in Table 3. The
temperature difference was taken as 319 F (60 F to -259 F) and 10 x 106 psi was used
as the modulus of elasticity for aluminum. Wall thicknesses in this table are based on
an allowable stress of 72 percent of specified minimum yield strength (SMYS). Loop
spacing is controlled by limiting total stress to the allowable values cited above.

TABLE 3. EXPANSION-LOOP SPACING FOR VARIOUS MATERIALS


34

The loop spacing U’ is limited by longitudinal stresses in the main line pipe out-
side the loops. The major longitudinal stress in this pipe is a friction stress between
the insulation and the pipe. This stress was computed conservatively by assuming that
the pipe was backfilled to a depth of 6 feet and that the insulation did not support any of
the earth load.

The effective length of added pipe assumes that each elbow causes a pressure drop
equivalent to 10 feet of straight pipe, In addition to loops with expansion chambers,
this design requires an intermediate anchor between each loop to equalize expansion.

In any vacuum-insulated line there is the additional problem of a dimensional


change imposed by the temperature difference between the inner pipe and the outer shell.
The connections between the two must be vacuum tight, locate the position of the inner
pipe with relation to the outer shell, and be flexible enough to follow the differential
movement.

Expansion Joints

In-line expansion joints of stainless steel bellow construction are available for
cryogenic service. Because the requirements of the cryogenic industry up to the
present time have been for low-pressure transfer lines, design and development of
bellows for high-pressure applications at liquid-methane temperatures have not been
extensive. Several manufacturers have indicated interest in the design and manufacture
of bellows for a pressure of 1000 psi; however, they also indicated that the present
state of the art would limit the axial movement of such a bellows to approximately
1 inch for either a lo-inch or a 20-inch-diameter pipe. The longitudinal contraction
for a temperature drop from 60 F to -259 F would be 107 inches/mile for 9 percent
nickel, 154 inches/mile for stainless steel, and 202 inches/mile for aluminum. Be-
cause of the large number of bellows required to compensate for the axial movement of
the LNG transfer line, the use of bellows does not appear to be economically sound at
the present time for a line pressure of 1000 psi.

For short distances (10-15 miles or less) where the LNG line is operating at or
below a pressure of 250 psi, the use of bellows may be justified due to the possible
greater maximum deflections of the expansion joints at this pressure. Reference 5 gives
a maximum axial deflection of 2.94 inches for a 6-convolution, Tube-Turns Series R
bellows joint operating at a pressure of 250 psi.

An LNG line using expansion joints only would have the bellows located in the line
back-to-back as shown schematically below:

where
x = intermediate anchor
= expansion joint.
Figure 16 shows details of a possible expansion joint, anchor, and concrete enclosure.
35

FIGURE 16. EXPANSION JOINT WITH CONCRETE ENCLOSURE

Each mile of an LNG line would require approximately 37 bellows for 9 percent
nickel steel, 53 bellows for stainless steel, and 69 bellows for aluminum, The ex-
pansion joints for cryogenic service are generally made from stainless steel. A
circumferential thermal stress develops at a welded joint of stainless steel and 9 per-
cent nickel due to the dissimilar contraction of the two metals when the temperature
drops from 60 F to -259 F. The joining of stainless steel to aluminum would have to be
accomplished by using a flanged joint. At this joint the two metals must be isolated
from each other in order to prevent galvanic corrosion.

Allowable Design Stress of Pipe

The code for gas-transmission lines, USASI B31,8, is not specifically applicable
to the pipe materials and operating temperatures of LNG pipelines. Allowable stresses
are given in the Petroleum Refinery Piping Code, USASI B31.3, which covers stain-
less, 3-1/2 percent nickel, and 6061-T6 aluminum piping. No allowable stress is given
for either 9 percent nickel or 5083-O aluminum piping, However, the code does give
an allowable stress of 30,000 psi for 9 percent nickel plate and 9750 psi for 5083-O
aluminum plate.

The ASME maximum allowable stresses for 9 percent nickel and 5083-O aluminum
are given as 23,750 psi* and 10,000 psi, respectively.

The allowable design stress in USASI B31.3 is based on room-temperature mate-


rial properties. However, because the tensile and yield strengths generally increase
with decreasing temperature, the design of an LNG line operating at -259 F will have an
additional safety factor above that generally included in a conventional gas-transmission
line.

The allowable design stresses as given in USASI B31.3 are compared with those
allowed under the current USASI B31.8 Code in Table 4. It should be pointed out that
*ASME Code Case 1308.
36
because USASI B31.3 deals primarily with refinery piping and does not account for in-
creased strength at low temperature, the use of this code for the design of an LNG line
is not directly applicable. It is not intended, in this report, to infer an allowable de-
sign stress for LNG pipelines; this is a matter requiring considerable thought and
study, The information given in Table 4, however, is considered useful in roughly
approximating required wall thicknesses for LNG lines. For the majority of the mate-
rials in Table 4, the USASI B31.3 allowable design stress corresponds to approximately
60 percent of SMYS. The only exception is 9 percent nickel which has an allowable
design stress equal to 50 percent of SMYS.

TABLE 4. ALLOWABLE DESIGN STRESSES

The allowable contraction stress for a temperature range from 100 F to -259 F
and 7000 or less temperature cycles is 1.5 times the basic stress value listed above.
If the longitudinal stresses caused by pressure, weight, or other sustained loadings
remain below the basic stress values listed above, the differences may be added to the
allowable expansion stress.

Table 5 gives the required wall thicknesses for the materials and the allowable
design stresses listed in Table 4 for a pipe having nominal pipe diameter of 10 inches
and subjected to a pressure of 1000 psi. The longitudinal joint factor, E, was taken as
unity.
37

TABLE 5. PRESSURE DESIGN THICKNESSES REQUIRED FOR A 10-INCH-DIAMETER PIPE SUBJECTED


TO A 1000 PSI PRESSURE ( a )
38

PIPE MATERIALS

Significant Properties

A major problem from a material standpoint in the construction of an LNG pipe-


line is the loss of ductility exhibited by some materials at low temperatures. One key
to the loss of ductility is found in the crystal lattice. Generally metals that have a
face-centered-cubic lattice such as copper and aluminum show no loss of ductility at
low temperatures. Members of the body-centered-cubic class such as iron will
generally fail with limited ductility below a certain transition-temperature range.
Austenitic stainless steels have a face-centered-cubic lattice and, therefore, have no
ductile-brittle transition temperature. In other ferrous alloys, the transition tempera-
ture is lowered by reducing carbon content or by increasing nickel content.

Notch toughness is of prime importance in low-temperature materials. Many


materials lose toughness and ductility as the temperature goes down. Although impact
properties are obtained mostly from Charpy Keyhole and V-notch tests, these tests
are at best indirect measurement of actual fracture toughness.

Ductility and toughness are sometimes used interchangeably, Actually, they are
distinctly different properties. A material that is ductile, as measured by elongation,
may have poor toughness, particularly at cryogenic temperatures. Conversely, some
high- strength steels may have low elongation and yet be tough in cryogenic
environments.

Tensile and yield strengths generally increase with lower temperature, but
tensile strength often increases more rapidly than yield strength.

Fatigue strength becomes important if service temperatures fluctuate between


ambient and ultralow levels, especially if stresses also vary. Fatigue- strength data
for many low-temperature and cryogenic materials are not available at present. How-
ever, for steels used in cryogenic shafting, pump components, and similar rapidly
operating equipment, standard room-temperature fatigue data are commonly used in
de sign. This allows an added safety factor, since fatigue strength increases as
temperature is reduced.

Corrosion resistance is not nearly the problem at cryogenic temperatures as it is


at ambient temperature, simply because low temperatures inhibit most chemical
reactions.

Changes in material characteristics can introduce new variables in the stress


computations at low temperatures. The modulus of elasticity increases somewhat at
low temperatures. The effect does not become serious because thermal expansion de-
creases with decreasing temperature. However, provision for pipe contraction in an
underground location requires special consideration.

In order to be suitable for the proposed service, a material must meet the follow-
ing engineering requirements:

(1) Exhibit an adequate yield strength


39

(2) Be ductile enough to withstand fabrication into pipe and field bending

(3) Possess adequate weldability for mill welding (if not seamless) and
field welding

(4) Possess suitable fracture-toughness properties at operating


temperatures.

Preliminary screening of data in approximately 40 publications indicated that


the materials discussed below might meet the requirements. The materials discussed
in the following sections are not the only materials which could be used for cryogenic
applications. Representative materials have been selected from various alloy groups
to provide some indication of the suitability of the group as a whole for the construction
of an LNG line.

Before presenting the results of the material survey, however, a discussion rela-
tive to fracture resistance will be presented because of its importance as a criteria for
material acceptibility.

Resistance to Fracture

Fracture-toughness properties are important from two viewpoints; propagation


resistance and initiation resistance, If a fracture should get started, propagation
resistance should keep it as short as possible. NG-18 research has shown that the
generalized determinant of whether a fracture will be long or short is the relation of
the fracture speed relative to the decompression velocity in the pressurizing medium. (6)
When a break occurs in a pipeline, pressure is not released from everywhere in the
line at once; it is exhausted from behind a decompression front which propagates
through the pressurizing medium at the acoustic velocity. If the fracture is faster than
the decompression velocity, the fracture cannot unload (the nominal stress at the
propagating fracture front remains the same as the original nominal stress before the
break) and the fracture tends to continue to propagate. On the other hand, if the frac-
ture is slower than the decompression velocity, it can unload and therefore would be
expected to arrest.

Steels in use for gas-transmission service have transition temperatures below


which they fail in a cleavage mode with speeds higher than the decompression velocity,
and above which they fail in a shear mode with speeds lower than the decompression
velocity, Charpy V-notch impact tests and the Battelle drop-weight tear test are cap-
able of identifying this transition temperature(7). The drop-weight tear test does this
directly with the temperature at which the broken-test specimen exhibits about 85 per-
cent shear area correlating with the full-scale transition temperature. The Charpy V-
notch test results may also be used but pipe-wall-thickness differences become a
factor - see Reference 7.

Some of the steels contemplated for LNG pipeline use will also have a transition
temperature below which they may be expected to fail in a relatively high-speed cleavage
fracture. Other steels contemplated for LNG pipeline use will not. In general, the
materials which do not have transition temperatures and which, therefore, always fail
in relatively slow-speed shear, are usually more expensive.
40

A crucial point, when considering materials which do have transition tempera-


tures for LNG pipeline use, lies in the relation between the fastest anticipated fracture
and the decompression velocity in the LNG. It may not be necessary to insist that the
transition temperature of the material be below the service temperature if it can be
demonstrated that the fastest anticipated fracture for a given material is less than the
LNG decompression velocity and therefore would be expected to be short. In the
material survey following, not enough of the right information is generally available to
ascertain this with certainty. Materials have been included which available impact data
indicate would be expected to have relatively high transition temperatures, keeping in
mind the possibility of “telescoping ‘I fracture-toughness properties.*

Van Itterbeek, et al. , (8) have calculated decompression in liquid methane based
on some experimentally determined thermodynamic data. The results of these calcula-
tions in the area of interest for LNG pipelines are presented in Table 6. They provide
a reasonable estimate for the decompression velocity in pressurized LNG. It is not
known whether the perturbations apparent are the result of insufficient thermodynamic
data, a possibility which the authors admit, or are a true representation of peculiar
characteristics of LNG in this region. Based on the fastest cleavage speeds of gas-
transmission-pipe steels** it may well be that the cleavage speeds of some
"economical” steels will be lower than the LNG decompression velocity, If this is true,
it allows consideration of many more materials than originally thought possible because
of the transition- temperature limitation. The survey of possible materials has been
made with this presumption in mind.

TABLE 6. ACOUSTIC VELOCITIES IN LIQUID METHANE

*A significant factor is that the LNG pipeline will have a pronounced temperature profile. Thus, it may be possible to “tele-
scope” fracture-toughness properties, using higher transition temperature materials at the warmer temperatures.
**The fastest cleavage speed measured in the NG -18 program has been about 2900 fps in X-52 pipeline steels. An interesting and
important point is that the cleavage speed, for a given material, appears to reach a plateau below the transition temperature
and to remain relatively constant for all lower temperatures.
41

While the above is of vital importance and opens the door for considering many
more materials, fracture toughness from an initiation point of view must also be con-
tended with. Current Pipeline Research Committee research is providing a means of
evaluating fracture toughness from an initiation point of view (9). This research is
demonstrating that a fracture-toughness index, Kc, can be obtained by conducting a
full-scale pressure test on a piece of pipe containing a defect. This fracture-toughness
index, Kc, has been shown to be temperature sensitive, but starting at a much lower
temperature than the propagation transition temperature referred to above. Thus, the
Kc index and its transition temperature are thought to be capable of providing an effec-
tive comparison of the relative resistance to fracture initiation between materials.
Unfortunately, information in the literature relative to fracture-initiation resistance is
meager and judgements on suitability of materials must be made on the basis of avail-
able impact data, Eventually, however, fracture-initiation resistance of the most pro-
mising materials should be obtained,

Austenitic Stainless Steel

Austenitic stainless steels meet all the requirements for low-temperature ser-
vice. These materials remain tough at liquid-methane temperatures, and their coef-
ficients of thermal expansion and thermal conductivity are lower than those of nonfer-
rous cryogenic structural materials such as aluminum. Most of these steels show a
slight increase in yield strength as temperature is lowered and a greater increase in
tensile strength.

Type 304 stainless steel is the most widely used metal for cryogenic piping be-
cause of strength, ease of fabrication, and excellent low-temperature properties.
Fittings and flanges are stocked in a wide range of sizes.

Stainless steels, Types 304L, 316, 321, and 347, are generally used in place of
Type 304 only where special corrosion problems exist. Low- temperature properties
are essentially the same as for Type 304. Because Type 304 stainless steel would be
adequate for the construction of an LNG transfer line, the discussion of mechanical
properties will deal primarily with this stainless steel.

Tensile properties of annealed and cold-worked Types 304 and 304L stainless
steel are presented in Figures 17 and 18(10). It can be seen from these figures that
the tensile and yield strengths increase at low temperatures - in contrast to the reduc-
tion in ductility of the annealed material. The percent reduction of the cross-sectional
area and the percent elongation of cold-worked Type 304 stainless steel appear to in-
crease initially and then decrease with decreasing temperatures. Although the tensile
and yield strengths increase with lower temperatures, design of low-temperature
equipment may be based on room-temperature tensile properties. Thus, an additional
built-in safety factor (based on strength alone) may exist,

Fatigue strength of cold-worked Type 304 stainless steel is given in Figure 19(10).
Since it is probable that an LNG transfer line will be subjected to less than 104 cycles
of stress (due to pressure or temperature) fatigue strength does not appear to be an
important parameter.
42

FIGURE 17. TENSILEPROPERTIES OF ANNEALED FIGURE 18. TENSILE PROPERTIES OF COLD-WORKED


TYPES 304 AND 304L STAINLESS TYPES 304 AND 304L STAINLESS
STEEL(10) STEEL(10)
43

FIGURE 19. RECIPROCATING BEAM FATIGUE STRENGTH OF


COLD- WORKED TYPE 304 STAINLESS STEEL ( 1 0 )

Table 7 gives the low-temperature impact properties of annealed Type 304 stain-
l e s s s t e e l ( 1 1 ) . Table 8 shows the relation of impact strength of Type 304 stainless steel
after low-temperature exposures up to 1 yeart 11), The data in this table indicate that
the steel is relatively insensitive to aging factors, In addition to the satisfactory
performance of the basic material, welded sections of Type 304 retain good impact
properties even though some reduction is noted as shown in Table 9(11). Izod and
Charpy Keyhole impact test results as a function of temperature are given in Figure
20(10). It is believed that this material will have good resistance against fracture
initiation and propagation,

Austenitic stainless steels such as Type 304 can be welded by a variety of pro-
cesses, i n c l u d i n g s h i e l d e d - a r c , g a s - m e t a l - a r c , g a s - t u n g s t e n - a r c , a n d s u b m e r g e d - a r c
methods, For shielded-arc welding, coated electrodes are available for most of the
steels. In addition, electrodes with slight modifications of these compositions are
available for special welding characteristics.

Thermal expansion is an important factor in material selection, In some cases,


thermal expansion might influence the decision to make the inner and outer pipes of a
v a c u u m - j a c k e t e d l i n e o f t h e s a m e m a t e r i a l . It is also an important factor in the selec-
tion of welding electrodes and can be a limiting factor in the choice of such an electrode.
The coefficients of thermal expansion for several materials are given in Figure 21.

Table 10 gives the thermal conductivity and some other selected properties of
Type 304 stainless steel. A high conductivity reduces thermal gradients during cool-
down but results in slightly higher heat gains through the piping when the line is in
service.

It is believed that of the austenitic stainless steels, Type 304 is the most appli-
cable. Available data has indicated adequate strength and, while appropriate fracture-
toughness data is not readily available, resistance to fracture initiation and propaga-
tion appears to be good.

The greatest limitation is cost. B a s i c - m a t e r i a l c o s t s m a y b e o n t h e o r d e r o f


eight to nine times the cost of ordinary carbon steel, e. g. , A53 steel. T h i s c o m p a r i s o n
44

TABLE 7. LOW-TEMPERATURE IMPACT PROPERTIES OF ANNEALED TYPE 304


STAINLESS STEEL (11)

TABLE 8. LOW-TEMPERATURE IMPACT PROPERTIES OF TYPE 304


STAINLESS STEEL AFTER PROLONGED EXPOSURE (11)

TABLE 9. LOW-TEMPERATURE CHARPY V-NOTCH IMPACT PROPERTIES OF


INERT-GAS-SHIELDED METAL-ARC WELD IN 3-1/2-INCH
TYPE 304 STAINLESS STEEL PLATE(11)
45

FIGURE 20. IMPACT PROPERTIES OF TYPE 304 STAINLESS STEEL (10)

FIGURE 21. COEFFICIENT OF LINEAR EXPANSION OF SEVERAL MATERIALS (11 )


46

does not account for any differences in allowable stress levels. See “Discussion” sec-
tion of this report for more details.

Aluminum Alloys

General Considerations

Commercially pure aluminum has a tensile strength of about 13,000 psi. Working
the metal can approximately double its strength. Much larger increases in strength
can be obtained by alloying aluminum with small percentages of other metals such as
manganese, silicone, copper, magnesium, or zinc. Like pure aluminum, the alloys
can also be made stronger by cold-working. Some alloys are further strengthened and
hardened by heat treatments; aluminum alloys having tensile strengths in the vicinity
of 100 ksi are presently available.

Like that of most metals, the strength of aluminum alloys under static or
repeated loading increases with decrease in temperature. Ultimate strengths generally
increase more rapidly than do yield strengths. Typical mechanical properties of some
aluminum alloys are listed in Table 11.

Reference 12 discusses the applicability of aluminum alloys in cryogenic environ-


ments. Most aluminum alloys used in such applications are weldable and many provide
relatively high strength (Table 12).

2000 Series

The 2000-series aluminum alloys are essentially A1-Cu alloys often containing
lesser amounts of magnesium, manganese, silicone, or lithium. Copper forms an
47

TABLE 11. TYPICAL TENSILE PROPERTIES OF ALUMINUM ALLOYS AT CRYOGENIC


TEMPERATURES (12)
48

TABLE 11. (Continued)


49

TABLE 12. ALUMINUM ALLOYS FOR CRYOGENIC APPLICATIONS(12)


50

intermetallic compound with aluminum (CuA1 2 ) and is the primary strengthening ingre-
dient. High strength is achieved by solution annealing at a temperature sufficiently
high to dissolve the CuA12 second phase (approximately 900 to 1000 F), followed by aging
at a lower temperature (approximately 200 to 400 F) to enable coherent precipitate
particles to form, Silicon and lithium additions permit the aging reaction to occur at
room temperature, although the highest strength is generally still obtained with
thermal-aging treatment.

Figure 22 illustrates the tensile properties of four 2000-series aluminum alloys


in the T6 (thermally aged) condition ( 1 3 ) . The 2020 alloy contains lithium and manganese
in addition to copper and is the strongest alloy in the group. However, its ductility is
lower than that of the other three alloys, especially at temperatures below -100 F. The
notched-unnotched tensile-strength ratio indicates that the 2020 alloy is also consider-
ably more susceptible to notch embrittlement.

The strength of the other three alloys decreases in the order 2014, 2219, and
2119. Ductilities are comparable and are in the vicinity of 10 percent elongation at
temperatures down to -450 F. The notched-unnotched strength ratio is less than 1.0
at room temperature and below. However, when this ratio is near 1.0 it indicates that
the notch strength was higher than the yield strength, and yielding occurred before
fracture. Thus, a lack of notch toughness may be critical for the 2020 alloy only.

Figure 23 illustrates the effect of variations in treatment on the tensile properties


(13)
of the 2219 alloy . Straining an age-hardenable alloy after solution annealing but
prior to aging generally results in the precipitation of finer, more evenly distributed
second-phase particles conducive to higher strengths. Thus, the 2219 alloy offers
slightly higher strength in the T87 condition (strained 7 percent prior to aging) than in
the T6 condition. Ductility is also somewhat higher in the T87 condition, probably due
to minimization of grain-boundary precipitation. The lower notched-unnotched strength
ratio of material in the T87 condition is largely due to the increased notch severity.

The effect of welding on the tensile properties of 2014 and 2219 alloys in the T6
(13)
condition is illustrated in Figure 24 . Welding significantly decreases the strength
of both alloys. This decrease is probably attributable to the destruction of the high-
strength “aged” structure in the weld metal.

Welding also reduces the ductility to fairly low values. This is usually the
result of a high degree of grain-boundary precipitation in the weld metal, in particular
the A1-Fe-Si intermetallic compound. A general coarsening of the structure of the
weld metal also contributes to lower ductilities.

3000 Series

Alloy 3003 has higher strength than pure aluminum, good formability, weldability,
and toughness . This alloy can be strengthened by strain hardening, but welding leaves
a heat-affected zone of lower strength. Consequently, these alloys are not used for
applications requiring a high strength-to-weight ratio. Also, allowable design stresses
may be expected to be quite low (3350 and 5000 psi for 3003-0 and 3003-H14, respec-
tively, as given in the USASI B31.3 Code). Thus, this alloy may be undesirable for
high-pressure service due to the additional problems involved in welding thick-walled
aluminum piping,
51

FIGURE 22. TENSILE PROPERTIES OF 2000 SERIES ALUMINUM ALLOYS


IN THE T6 CONDITION (13)

FIGURE 23. EFFECT OF HEAT TREATMENT ON 2219 ALUMINUM ALLOY ( 1 3 )


52

5000 Series
Magnesium is the major alloying constituent in the 5000-series aluminum alloys,
Lesser quantities of manganese are generally included. Although an A1-Mg intermetal-
1ic compound can be formed, its precipitation during an aging treatment does not
increase the strength. A combination of cold work and solid-solution strengthening is
relied upon to attain improved strengths.

Figure 25 shows the minimum tensile strengths of some 5000-series aluminum


alloys in the annealed condition (14). Alloys 3003-0 and 6061-T6 are included in this
figure for comparative purposes,

Table 13 lists the results of tensile tests of some 5000 series alloys at several
(15)
temperatures . Figure 26 illustrates the tensile properties of three 5000-series
aluminum alloys in the H3X* condition (cold worked* and stabilized). (13) As a group
these alloys are characterized by low strength and high ductility in comparison with
the 2000-series alloys. The elongation of 5052-H32 and 5086-H34 alloys at low temper-
atures appears to be higher than at room temperature.

The notched-unnotched strength ratio of all of these alloys is slightly less than
1.0 at room temperature and decreases somewhat at low temperatures. However,
because of the excellent ductility at low temperatures and the spread between the yield
strengths and tensile strengths, notch embrittlement is probably not a problem for the
5000-series alloys at low temperatures.

Figures 27, 28, and 29 compare the welded and base-metal tensile properties of
the 5000-series aluminum alloys ( 1 3 ) . Welding produces a loss in strength but not as
much as was observed with the 2000-series alloys. This is no doubt attributable to the
different strengthening behavior of the alloys in the two series. The second-phase
distribution in the 2000-series alloys is highly sensitive to thermal exposure. Con-
sequently, its distribution is altered by welding, and there is a significant decrease in
strength. By virture of their reliance on solid solution as a strengthening mechanism,
the strength of the 5000-series alloys is not greatly affected by welding.

The effect of welding on the ductility of the 5000-series alloys is more pronounced
than its effect on strength. Nevertheless, the ductility is not reduced as much as was
noted for the 2000-series alloys. Similar to the 2000-series alloys, the reduction in
ductility is probably largely due to the formation of second-phase particles in grain
boundaries of the weld metal during solidification. The presence of impurities, mainly
iron, is responsible for the second-phase formation.

Figures 30, 31, and 32 present the Charpy V-notch impact-energy curves for
5052-F, 5086-0, and 5154-0 aluminum alloys ( 1 6 ) . These materials may be expected
to have good fracture resistance against propagation, but their resistance to initiation
is uncertain.

6000 Series

Toughness of the 6000-series alloys is slightly higher at low temperatures than


at room temperatures. Alloy 6061 is the oldest of this easily welded group and is one
of the most widely used; it is available in almost all product forms, These alloys are
*X = percent cold worked.
53

FIGURE 24. EFFECT OF WELDING ON THE TENSILE PROPERTIES OF 2014 AND


2219 ALUMINUM ALLOYS IN THE T6 CONDITION (13)

FIGURE 25. MINIMUM TENSILE STRENGTHS OF SOME ANNEALED


5000-SERIES ALUMINUM ALLOYS ( 1 4 )
55

FIGURE 26. TENSILE PROPERTIES OF 5000 SERIES FIGURE 28. EFFECT OF WELDING ON THE TENSILE
ALUMINUM ALLOYS IN THE H3X PROPERTIES OF 5086 ALUMINUM
CONDITION(13) ALLOY IN THE H34 CONDITION (13)

FIGURE 27. EFFECT OF WELDING ON THE TENSILE FIGURE 29. EFFECT OF WELDING ON THE TENSILE
PROPERTIES OF 5052 ALUMINUM PROPERTIES OF 5456 ALUMINUM
ALLOY IN THE H32 CONDITION(13) ALLOY IN THE H343 CONDITION ( 1 3 )
56

FIGURE 30. CHARPY V-NOTCH IMPACT ENERGY OF 5052-F ALUMINUM ( 1 6 )

FIGURE 31. CHARPY V-NOTCH IMPACT ENERGY OF 5086-O ALUMINUM (16)

FIGURE 32. CHARPY V-NOTCH IMPACT ENERGY OF 5154-O ALUMINUM (16)


57

extrudable, with 6063 being lower in cost and in strength than 6061. Alloys 6066
(forgings and extrusions), 6070 (extrusions), and 6071 (sheet and plate) have higher
static strengths and slightly higher weld strengths, but lower toughness than 6061 and
6063. In the as-welded condition, strengths of the 6000-series alloys are less than
those of the 5000 series. However, postweld aging or heat treatment and aging brings
the strengths to the upper limits of the range available in the 5000-series alloys.

Figures 33 and 34 present the tensile properties of 6061 aluminum alloy(16,17,18)


The Charpy keyhole impact energy for this alloy in the T6 condition is given in Figure
35. Although fractures in this material - if they should get started - would be expected
to be shear, the low energy level causes some concern. Also, fracture-initiation
resistance is uncertain.

7000 Series

The 7000-series aluminum alloys contain magnesium and zinc as major alloying
constituents. Lesser amounts of chromium, copper, and manganese are generally in-
cluded, Like the 2000-series alloys, the 7000-series alloys are dispersion strengthened.
High-strength results from precipitation of a Zn-Mg intermetallic compound during a
thermal aging treatment. Copper joins in the compound formation and is added as a
secondary strengthener. Chromium and manganese improve the resistance to stress
corrosion.

Figure 36 illustrates the tensile properties of four 7000 series aluminum alloys
in the T6 condition. (13) The alloys in this group provide higher strengths than alloys in
the 5000 series. Strength for the four alloys under consideration increases in the order
7002, 7079, 7075, and 7178, which is the order of increasing zinc, and copper plus
magnesium content. A corresponding decrease in ductility accompanies the increase in
strength. Alloy 7002 offers considerably more ductility than the other three alloys
and, in fact, is more ductile than any of the 2000 series alloys, Thus, 7002 provides
a good combination of high strength and superior ductility.

The indicated resistance to notch embrittlement of 7002 is better than that of any
other aluminum alloys evaluated. Notch embrittlement may be a problem with 7079,
7075, and 7178 alloys at cryogenic temperatures,

The effect of welding on the tensile properties of 7002 in the T6 condition is


illustrated in Figure 37. As for the 2000-series alloys, both the strength and ductility
are markedly reduced by welding. The explanation of these phenomena presented for
the 2000-series alloys also applies to the 7000-series alloys. It is worth noting, how-
ever, that the ductility of welded 7002 is somewhat superior to that of the welded 2014
and 2219 alloys at temperatures down to at least -200 F.

Reference 19 summarizes the developed experimental data on aluminum alloy


7039. Its combination of good tensile and yield strengths at low temperature and
adequate strength and ductility of naturally aged weldments at -320 F should make it
well suited to cryogenic application, However, very little impact-energy data are
available for this material, as well as for most of the other aluminum alloys,

The tensile properties of 0.75-inch 7039-T6 plate as given in Figure 38 indicate


that the elongation increases to about 15 percent at -320 F. Tensile-test data on sheet
58

FIGURE 33. TENSILE PROPERTIES OF 6061 ALUMINUM ALLOY ( 1 6 , 1 7 , 1 8 )

FIGURE 34. ELONGATION OF 6061 ALUMINUM ALLOY ( 1 6 , 1 7 , 1 8 )

FIGURE 35. CHARPY KEYHOLE IMPACT ENERGY OF


6061-T6 ALUMINUM ALLOY ( 1 8 )
59

FIGURE 36. TENSILE PROPERTIES OF 7000 SERIES ALUMINUM ALLOYS


IN THE T6 CONDITION (13)

FIGURE 37. EFFECT OF WELDING ON THE TENSILE PROPERTIES OF


7002 ALUMINUM ALLOY IN THE T6 CONDITION (13)
60

FIGURE 38. TENSILE PROPERTIES OF 0.7 5-INCH ALUMINUM ALLOY


7039-T6 PLATE, FROM 75 TO -423 F (19)
61

and plate thicknesses are shown in Table 14 and, for the most part, reflect an average
determined from tests on at least three specimens (19) . The notched tensile tests as
given in Table 14 are for specimens with a notch factor Kt = 6.3.

TABLE 14. TENSILE PROPERTIES OF 7039 ALUMINUM ALLOY PARENT METAL AT 75 AND -320 F(19)

The tensile-test results of representative welds, “as-welded”, are shown in


Table 15. The data for thicknesses up to 0.75 inch indicate that “bead off” (weld bead
machined off to conform to a round reduced section-type tensile specimen) tensile
strengths of about 51,000 to 53,000 psi can be attained after 15 to 30 days aging at room
temperature, using 7039 plate and X5039 filler alloy, without the necessity of postweld
heat treatment.

Weldability

The melting point of aluminum is lower than the temperature at which it glows.
Thus, a welder cannot use color as an indication of temperature and usually must be
experienced before being competent in welding this metal.

The high thermal conductivity requires high heat input for fusion welding. Heavy
sections may require preheating, Aluminum and its alloys develop a refractory oxide
film which influences welding characteristics. The oxide film must be broken up during
welding if the liquid filler metal is to flow and wet properly. In practice, this film is
removed either by fluxes or by the action of a welding arc in the inert-gas atmosphere.
The latter is preferred, because fluxes capable of performing this task are corrosive
if left on the joint.
63

All nonheat-treatable aluminum alloys are work hardenable and are welded by the
gas metal-arc and gas tungsten-arc processes. The welding operation anneals local
areas, however. Thus, the weld zone exhibits strength characteristics comparable to
the annealed condition rather than the cold-worked condition.

Strength of heat-treated aluminum alloys is decreased by overaging or re-solution


treating resulting from the heat of welding. However, since full annealing requires
considerable time, properties in the heat-affected zone are those of a partially annealed
material. These properties are usually lower than the strengths and elongation of
nonheat-treatable alloys of the 5000 series after welding, These alloys can be heat
treated after welding to restore properties if a heat-treatable filler metal has been
used.

Aluminum-alloy 6061 can be used in the as-welded condition or it can be postweld


heat treated. Higher as-welded strength is obtained by rapid welding. A further in-
crease in strength is achieved by welding in the solution heat-treated condition, then
postweld aging. This technique is more successful in thin gages than in heavy gages,

Of all weldable aluminum alloys, the 2000 series can be aged to develop the
highest parent-metal strength, but several 5000 and 7000 series alloys have higher
weld strength in the as-welded condition.

Joining aluminum and stainless steel, e. g. , stainless steel bellows in an aluminum


line, will present some difficulties. Reference 20 discusses the possibility of joining
aluminum to stainless steel by brazing and by diffusion bonding.

Aluminum-alloy 6061 was brazed successfully to 304L and 321 stainless steel.
Joints made by this process have withstood vibration, cryogenic thermal shocks, and
pressure burst tests.

Diffusion bonding will bond 2219 and other aluminum alloys to 321 stainless steel.
The idea is to incorporate the joint in a separate part, make the bond, and then weld
the several metals of the part to their counterpart metal of the larger assembly.

Thermal Properties

The instantaneous coefficients of linear expansion of some aluminum alloys are


( 1 5 )
given in Figure 39 . The higher-strength A1-Mg alloys should fall in the range be-
tween alloys 5052 and 5056.

Values of specific heat for most of the aluminum alloys fall within the width of the
line in Figure 40 ( 1 5 ) .

Thermal conductivities of some aluminum alloys are illustrated in Figure 41. (15)
The A1-Mg alloys should have thermal conductivities approximately the same as alloy
5154.

Recommendations for LNG Piping

Aluminum alloys are considered good candidates for use in an LNG pipeline. It
appears that good strength can be achieved and resistance to fracture propagation
64

FIGURE 39. COEFFICIENTS OF EXPANSION OF SOME ALUMINUM

(15)
FIGURE 40. SPECIFIC HEAT OF SOME ALUMINUM ALLOYS

FIGURE 41. THERMAL CONDUCTIVITIES OF SOME ALUMINUM ALLOYS ( 1 5 )


65

appears satisfactory for several alloys. Fracture-initiation resistance, however, is


uncertain and will have to be determined,

Of the alloys discussed above, 5456-0, 5083-0, 6061-T6, and 7002-T6 appear to
be the most: promising,

The possible use of aluminum is attractive from an economical viewpoint consider-


ing base costs of material alone (i. e. , neglecting any imposed costs due to design con-
siderations - large thermal contraction compared to steel, for example), Aluminum
base-material costs may be expected to be on the order of three to four times the base
cost of an ordinary carbon steel. However, this does not acknowledge any difference
in allowable stress levels. See “Discussion” section of this report for more details.

Nickel-Base Alloys

Nickel-base alloys such as Monel and Inconel have been developed primarily for
applications requiring corrosion resistance and high-temperature strength. However,
by virtue of having a fcc crystal structure, many nickel alloys display attractive
properties at low temperatures as well.

Chromium and copper additions serve as solid-solution strengtheners, but most


nickel alloys rely to a large extent on the precipitation of a second phase during an
aging treatment to achieve maximum strength. The aged condition provides higher
strength at some sacrifice in ductility. Nevertheless, no tendency toward brittleness
has been reported at temperatures down to -423 F.

Representative chemical analyses, expressed in percent, for Monel and Inconel,


are as follows:

Figures 42 through 45 present the tensile properties of Monel and Inconel(21).


For a temperature decrease from 80 to -259 F, the tensile strength of Monel in the
annealed condition increased from 75 to 106 ksi. The elongation of annealed Monel
decreased from 50 to 45 percent for a temperature drop from 80 F to -150 F and then
increased to 55 percent at -259 F.

Inconel (cold drawn 10 percent) also exhibited an increase in tensile strength,


yield strength, and elongation with a decrease in temperature. The elongation for
Inconel (cold drawn 50 percent) remained approximately constant at 10 percent for a
temperature range from -80 to -320 F.

Figure 46 presents Charpy V-notch impact energy for Monel as a function of


(21)
temperature . This material exhibited excellent fracture-toughness properties in
both the annealed and hot-rolled condition.
66

FIGURE 42. STRENGTH OF MONEL ( 2 1 ) FIGURE 43. ELONGATION OF MONEL ( 2 1 )

FIGURE 44. STRENGTH OF INCONEL ( 2 1 ) FIGURE 45. ELONGATION OF INCONEL ( 2 1 )


67

For Inconel the Charpy V-notch impact energy (Figure 47) is considerably higher
for the hot-rolled condition than for cold-drawn (50 percent) material. However, the
impact energy for cold-drawn (50 percent) Inconel appears to be constant at approxi-
mately 60 ft-lb, as the temperature decreases from 80 F to -320 F.

Both of these materials would be excellent candidates for an LNG pipeline.


However, they are very expensive - even more costly than Invar (36 percent nickel)
which has a base-material cost of about 10 to 12 times that of ordinary carbon steel.
Unless overall costs are significantly lowered by the fact that expansion bellows or
loops are not needed, these materials do not appear to be likely candidates.

Nickel Steels

Of the more common alloying elements, nickel has been shown to be outstanding
for lowering the transition temperature of ferritic steels. The general effect of nickel
content on the impact-energy value is shown as a function of temperature in Figure
48. (22) Also, the coefficient of thermal expansion decreases in direct proportion to
the nickel content in the 2-1/4 to 9 percent nickel steels. Invar, a 36 percent nickel-
alloy steel, has an expansion coefficient so low that thermal-contraction stresses are
relatively insignificant, and the use of bellows or expansion joints is unnecessary.

3-1/2 Percent Nickel Steel

Low-carbon 3-1/2 percent nickel steel has had wide use in land-based facilities
for the storage of liquefied gases at temperatures down to and slightly below -150 F. I t
is frequently specified for tanks, vessels, and piping in order to handle liquid propane,
carbon dioxide, and liquid ethylene.

The minimum tensile and yield strengths as given in USASI B31.3 for 3-1/2 per-
cent nickel steel Grade D are 65,000 and 37,000 psi. The corresponding stresses for
Grade E are 70,000 and 40,000 psi, respectively.

Effects of heat treatment and plate thickness are shown in Figure 49(11). This
steel contained 0.13 percent carbon, 0.50 percent manganese, 0.29 percent silicon,
and 3.45 percent nickel. Figure 49 indicates that quenching and tempering of 1/2-inch
plate results in better fracture toughness than those obtained from either normalizing
and tempering or from normalizing only.

Normally expected Charpy impact properties of normalized and tempered 1/2-


inch-thick, 3-1/2 percent nickel steel plate are shown in Figures 50 and 51 (11) . The
keyhole-notch Charpy values shown in Figure 50 are plotted as a scatterband. A curve
drawn within the band indicates the median values or normal expectancy at various
temperatures. Figure 51 shows a similar curve for Charpy V-notch test results.
Table 16 shows the notch-toughness properties of 3-1/2 percent nickel steel at - 150 F
for half-inch welded plate ( 1 1 ) .

Selected thermal properties for 3-1/2 percent nickel steel are given in Table 17,
68

FIGURE 46. CHARPY V-NOTCH IMPACT ENERGY OF MONEL (21)

FIGURE 47. CHARPY V-NOTCH IMPACT ENERGY OF INCONEL (21)


69

FIGURE 48. HOW NICKEL AFFECTS KEYHOLE-NOTCH CHARPY


IMPACT TEST PROPERTIES ( 2 2 )

FIGURE 49. EFFECT OF HEAT TREATMENT AND PLATE THICKNESS ON


KEYHOLE-NOTCH IMPACT PROPERTIES OF
3-1/2 PERCENT NICKEL STEEL PLATE ( 1 1 )
70

FIGURE 50. CHARPY KEYHOLE-NOTCH IMPACT VALUES FOR


3-1/2 PERCENT NICKEL-STEEL PLATE ( 1 1 )

FIGURE 51. CHARPY V-NOTCH IMPACT VALUES FOR 3-1/2 PERCENT


NICKEL-STEEL PLATE(11)
71

TABLE 16. NOTCH-TOUGHNESS PROPERTIES OF 3-1/2 PERCENT NICKEL-STEEL


WELDED PLATE AT -150 F ( 1 1 )

For manual metal-arc welding, low-hydrogen-type electrodes will deposit weld


metal containing not over 0.10 percent carbon and 3 to 3-1/2 percent nickel. Welding
experience using the submerged-arc and inert-gas metal-arc processes has been satis-
factory, but limited, for nickel steel plate chiefly because adequate supplies of nickel-
steel wire have only recently become available. One process which has been used
successfully with submerged-arc equipment consists of using low-carbon steel wire as
filler metal and obtaining the desired nickel content from the flux. Results to date
indicate that multipass techniques will be required to meet low-temperature test
requirements in the weld metal.

9 Percent Nickel Steel

Nine percent nickel steel which was developed for use in the -150 to -320 F
temperature range offers several advantages:
Probable high allowable design stress
Moderate cost
72

High tensile and yield strengths


Good notch toughness at low temperatures
Good weldability.

Additional interest in this steel has resulted from the special ruling in ASME Code Case
1308 which permits the use of 9 percent nickel steel without postfabrication stress
relief.

Until recently, 9 percent nickel steel was available only in plate form and, there-
fore, found little application in piping installations. Now, however, it is available in
the most common seamless pipe sizes, welding fittings, and flanges, as well as hot-
rolled and cold-drawn seamless tubing up to 10-3/4-inch outside diameter.

The relation of tensile properties to temperature for base metal and for as-welded
specimens are presented in Figures 52 and 53, respectively(23) . The tensile- and
yield-strength curves in Figure 53 represent values for the weld metal since all frac-
tures occurred at the welds.

Yield strength of the base metal increased from 100,000 psi at room temperature
to 136,000 psi at -320 F, and the corresponding values for the welded specimens were
60,000 and 80,000 psi. The tensile strength at -320 F was 41 percent greater than
that measured at room temperature (191,000 to 135,000 psi) for the base-metal speci-
mens. The welded specimens had tensile strengths of 140,000 psi and 96,000 psi at
-320 F and at 70 F, respectively, In both Figures 52 and 53 the elongation appears to
increase slightly with decreasing temperature. The reduction of area for base-metal
specimens dropped slightly from 61.5 percent at room temperature to 53.0 percent at
-320 F. The corresponding figures for welded specimens were 22.0 and 40.0 percent.
The largest increase in the reduction of area values occurred between -100 F and -200
F. Reasons for this anomaly, i, e., increase in reduction in area with decrease in
temperature, are not known.

The heat treatment recommended in ASTM A-353 is double normalizing at 1650 F


and at 1450 F, followed by stress relieving at 1025 to 1085 F. Nine percent nickel
steel is also available from the mill in the quenched and tempered condition. The im-
pact properties of quenched and tempered steel are better than that of double-normalized
and tempered (or stress relieved) over the range from room temperature to -320 F.
Effect of heat treatment and plate thickness on keyhole-notch Charpy impact properties
of 9 percent nickel steel is given in Figure 54(34).

Acceptance of 9 percent nickel steel in the quenched and tempered condition by


the ASME Boiler and Pressure Vessel Committee (Case 1308-5) makes stress relief of
finished vessels no longer necessary whenever the walls are 2 inches and under in
thickness.

Keyhole-notch Charpy values for 9 percent nickel-steel plate in thicknesses up to


(11)
and including 1 inch, double normalized, and tempered, are shown in Figure 55
Because there is no inflection point on the keyhole-notch Charpy curve for this steel
at temperatures down to -320 F, tests are usually not made at lower temperatures.

Charpy V-notch impact test data are given in Figure 56(25), At -320 F there does
not appear to be a significant difference in the amount of energy absorbed by the
73

FIGURE 52. RELATION OF TENSILE PROPERTIES TO TEMPERATURE


FOR NORMALIZED AND STRESS-RELIEVED 9 PERCENT
NICKEL-STEEL PLATE(23)

FIGURE 53. RELATION OF TENSILE PROPERTIES TO TEMPERATURE


FOR 9 PERCENT NICKEL-STEEL PLATE WELDED WITH
INCONEL-TYPE WIRE AND TESTED AS WELDED (23)
74

FIGURE 54. EFFECT OF HEAT TREATMENT AND PLATE THICKNESS ON


KEYHOLE-NOTCH CHARPY IMPACT PROPERTIES OF
9 PERCENT NICKEL STEEL ( 2 4 )

FIGURE 55. CHARPY KEYHOLE-NOTCH IMPACT VALUES FOR


9 PERCENT NICKEL-STEEL PLATE ( 1 1 )
75

quenched and tempered, and double-normalized and tempered transverse specimens,


However, at room temperature the double-normalized and tempered transverse speci-
mens have a slightly higher maximum-energy plateau.

Drop-weight tests (used to determine nil-ductility temperatures (NDT)) have been


performed on both 1/2 and 1-inch-thick plate, and fracture did not propagate at test
(11)
temperatures as low as - 320 F , Thus , in these tests, the NDT was at least as low
as -320 F, possibly indicating good resistance to fracture initiation,

Charpy keyhole- and V-notch impact tests have been made on welded specimens
with the notch in the weld metal and in the heat-affected zone(23). Results of impact
tests are given in Tables 18 and 19. Maximum and minimum values are shown for each
test which usually consisted of breaking three specimens. In both Tables 18 and 19, the
Charpy V-notch values at -320 F as well as at room temperature are higher than those
for the Charpy keyhole-notch specimens.

Fatigue properties for 9 percent nickel steel are shown in Figure 57(23). Endur-
ance limits of 72,000 psi and 27,000 psi were obtained on unnotched and notched speci-
mens. The endurance limit of 27,000 psi obtained with notched specimens indicates a
notch sensitivity of 2.66.

TABLE 18. CHARPY IMPACT TEST RESULTS ON AS-WELDED 5/16-


INCH-THICK 9 PERCENT NICKEL-STEEL PLATE(23)

Selected thermal properties of 9 percent nickel steel are given in Table 20,
Because the coefficients of expansion are lower for 9 percent nickel steel than for the
aluminum alloys, smaller dimensional changes will result in the pipeline for a temper-
ature decrease from ambient to -259 F. This will result in the use of fewer expansion
joints or expansion loops to compensate for the longitudinal contraction of the line.
76

FIGURE 56. RESULTS OF 3/4-WIDTH CHARPY V-NOTCH IMPACT TESTS


ON 3/8-INCH-THICK PLATES OF 9 PERCENT
NICKEL STEEL ( 2 5 )

FIGURE 57. S-N CURVES FOR NOTCHED AND SMOOTH ROTATING BEAM
FATIGUE SPECIMENS OF NORMALIZED AND STRESS-
RELIEVED 9 PERCENT NICKEL-STEEL, ROOM-
TEMPERATURE TESTS ( 2 3 )
77

TABLE 19. CHARPY IMPACT TEST RESULTS ON AS-WELDED 1/2-


INCH-THICK 9 PERCENT NICKEL-STEEL PLATE(23)

TABLE 20. SELECTED THERMAL PROPERTIES OF 9 PERCENT


NICKEL STEEL ( 1 1 )
78

lnvar

Invar is an alloy containing nominally 36 percent nickel and the balance iron. I t
is an austenitic alloy having excellent ductility and impact energy over extremely low
temperatures. Its outstanding and most useful property is its low coefficient of thermal
expansion from temperatures of approximately -400 F to approximately 600 F.

Recently, the advantages of Invar’s low coefficient of expansion and excellent


ductility and impact energy at low temperatures were recognized by the cryogenic
industry for the construction of vacuum-jacketed piping systems. Welded Invar pipe
has been used in sizes up to 6 inches in vacuum-jacketed transfer lines.

Invar cannot be strengthened by heat treatment and obtains only moderate in-
creases in strength by cold working. Combined with a relatively low work-hardening
rate, the extremely good ductility and toughness of this alloy have made drawing reduc-
tions in excess of 40 percent possible without intermediate annealing. Annealing is
normally accomplished by heating at 1450 F for 30 minutes per inch of thickness; how-
ever, progressive softening occurs at temperatures above 1000 F. Grain growth be-
comes apparent at and above 1900 F, causing a slight decrease in strength.

The room-temperature mechanical properties and some physical properties of


Invar are given in table 21 (26) . The cryogenic mechanical properties of this alloy are
p r e s e n t e d i n F i g u r e 5 8 (26).

TABLE 21. TYPICAL AMBIENT-TEMPERATURE PROPERTIES


OF ANNEALED INVAR ( 2 6 )

The stress-strain diagram for Invar, as a function of temperature, and the varia-
tion of the modulus of elasticity with temperature are given in Figures 59 and 60(27).
It is evident from Figure 59 that the yield strength increases from 94 ksi at 70 F to
approximately 140 ksi at -320 F. The modulus of elasticity decreases slightly with
temperature in Figure 60.

Prior to the recent development of suitable welding procedures and alloys, Invar
was not often used for structural applications. In order to overcome the gross defects
that occurred when attempts were made to weld Invar, selenium (an element added to
improve machinability) was eliminated and other residual elements were controlled with
titanium additions.
79

FIGURE 58. TYPICAL MECHANICAL PROPERTIES OF INVAR PLATE ( 2 6 )


80

FIGURE 59. STRESS-STRAIN DIAGRAM FOR INVAR (27)

FIGURE 60. MODULUS OF ELASTICITY OF INVAR ( 2 7 )


81

A comparison of mechanical properties developed from test plates gas-tungsten-


arc welded with various filler metals of composition similar to the base metal is shown
in Table 22(2 6 ) . The compositions of the materials tested are shown in Table 23(26).
Note the erratic ductility of welds not containing titanium, especially at cryogenic
temperatures.

AS seen from the data in Table 22, weld joints made with titanium-bearing filler
metal developed consistent ductility of adequate magnitude. Although postweld annealing
improved both ductility and impact strength, as-welded properties are adequate for
most applications. The increased manganese content and addition of about 1 percent
titanium to the modified filler metal did not appreciably affect the expansion character-
istics of the deposited weld metal,

The compatibility of an alloy with other materials is a limiting factor in its use
for widespread application. Invar, while possessing unique properties of low thermal
expansion, also has the added advantage of weld compatibility with nickel-base alloys
and austenitic stainless steels.

Recommendations for LNG Pipeline

Of the nickel steels, possibly the most applicable for LNG pipelines would be the
9 percent nickel steels in the quenched and tempered condition, This is primarily
based on strength, toughness, and cost considerations, Invar is perhaps better from
the fracture-toughness viewpoint. Invar, aside from its higher base-material costs
(10-12 times that for carbon steel) would have lower installation costs because expan-
sion loops or bellows would not be required. The 3-1/2 percent nickel steel, while it
does not have the fracture-toughness levels that 9 percent nickel does, is significantly
less expensive. Base-material costs for 3-1/2 percent and 9 percent nickel steels are
about three to three and one half times and six to seven times those for ordinary carbon
Steel, respectively. The 3-1/2 percent nickel steel may be qualified for LNG pipeline
use by additional fracture-toughness testing.

Quenched and Tempered Steels

Some of the present quenched and tempered steels may be suitable for LNG pipe-
line service if it can be demonstrated that

(1) their fracture speeds are sufficiently below the decompression velocity
in LNG

(2) their resistance to fracture initiation is sufficient.

Examples are 5 Ni-Cr-Mo-V and HY-80.

Typical stress-strain curves for HY-80 steel and 5 Ni-Cr-Mo-V steel are shown
in Figure 61 ( 2 8 ) , and typical mechanical properties are presented in Table 24 (28) .

Typical Charpy V-notch impact results for HY-80 and 5 Ni-Cr-Mo-V steels as
illustrated in Figure 62 show that fracture transition occurs at a slightly higher
82

TABLE 22. EFFECT OF TITANIUM-MODIFIED FILLER METAL ON THE MECHANICAL PROPERTIES OF


GAS-TUNGSTEN-ARC WELDED INVAR ALLOY PLATE 1/4-INCH THICK(26)

TABLE 23. PERCENTAGE COMPOSITION OF INVAR BASE METAL AND WELD


FILLER METAL (26)
83

FIGURE 61. STRESS-STRAIN CURVES FOR HY-80 AND 5Ni-Cr-Mo-V STEEL (28)

FIGURE 62. CHARPY V-NOTCH IMPACT TEST RESULTS FOR HY-80


AND 5Ni-Cr-Mo-V STEELS(28)
84

(28)
temperature for 5 Ni-Gr-Mo-V steel than for HY-80 steel . The NDT values for
(28)
HY-80 steel and 5 Ni-Cr-Mo-V steel are - 160 F and - 120 F, respectlvely .

TABLE 24. TYPICAL MECHANICAL (PROPERTIES OF KY-80 STEEL AND


5 Ni-Cr-Mo-V STEEL 2 8 )

In general, these steels can be welded by any of the usual production methods.
Preheating is not usually required unless welding is to be done in cold weather. In
fact, preheating may be harmful because it may prevent the desired level of mechanical
properties from developing in the heat-affected zone.

The suitability of these materials will have to be proven by further test for use as
an LNG pipeline material. In view of the possible economic advantage, this is con-
sidered worthwhile.
85

INSULATION MATERIALS

There are many materials available which can be used for low-temperature insula-
tion. The majority of these materials depend on finely divided air spaces to retard the
transfer of heat.

The insulating materials can be categorized as belonging generally to the following


four classes:

(1) Fibrous
(2) Cellular
(3) Granular
(4) Reflective.

The fibrous insulation such as hair felt, asbestos, and mineral wool offer little resis-
tance to the passage of water vapor through the air spaces, have a relatively high
thermal conductivity, and are not recommended for LNG pipeline usage. The granular
and reflective materials will be discussed in the sections dealing with powder insulations
and multilayer insulations.

Cellular Insulation

The cellular insulations such as foamed glass, plastic foams, cork, and balsa
wood are discussed presently. The following properties of foams are desirable for
monolithic insulation systems exposed to LNG temperatures:

(1) Low tensile and compressive moduli


(2) Low thermal- expansion coefficient
(3) Low Poisson’s ratio
(4) Low thermal conductivity
(5) Low density.

It is desired that the foamed insulations provide good resistance to heat transfer
under the imposed temperature boundaries, withstand the maximum internal stress
generated by mechanical and thermal forces, and resist deformations in a radial di-
rection due to earth loads. The following major types of foams appear to be best suited
for an LNG underground pipeline: urethane, polystyrene, epoxy, silicone, phenolic and
syntactic.

Urethane Foams

Urethane foams have high strength, good abrasion and chemical resistance, and
can be cast in slabs, molded, sprayed, or foamed-in-place. They equal or exceed most
of the physical characteristics of the other foams and have the widest range of formula-
3
tions. Densities range from 1,5 to 70 lb/ft . Urethane has the lowest thermal con-
ductivity (k-factor) of all the foams under consideration.
86

The density of rigid urethane foams affects most properties. The mechanical
properties such as tensile and compressive strengths increase with density as indicated
in Figure 63 ( 2 9 ) . However, with increasing density, the k-factor (coefficient of con-
ductivity) also increases (Figure 64) with the disadvantageous effect of greater heat influx
through the insulation walls. The best choice for a particular application from thermal
considerations will be that density which just satisfies the tensile- and compressive-
strength requirements.

Figure 65 gives the relationship between temperature and thermal conductivity for
various materials ( 3 0 ) . It is evident from Figure 65 that urethane in the 2 lb/ft3 formu-
lation has the lowest thermal conductivity of all the foams under present consideration.
However, at a mean temperature of -100 F, the thermal conductivity of all these insula-
tions at their optimum formulations ranges from 0.12-0. 20 Btu-in./(hr)(ft 2 )(F). The one
exception is cellular glass, which due to its high density of 9 lb/ft3 has a k-factor of
0.29 at a mean temperature of -100 F.

Figure 66 presents tensile , shear, and compressive strengths of ‘urethane foam


as a function of temperature ( 3 1 , 3 2 ) . For a density of 2 lb/ft3 and a temperature drop
from 75 to -259 F, the tensile and shear strengths decreased from 72 to 58 psi and from
3
47 to 44 psi. The compressive strength of urethane at 3.4 lb/ft increased steadily from
39 psi at room temperature to 124 psi at -259 F.

There is very little information available on the behavior of foams subjected to


compressive loads at LNG temperatures, primarily because the majority of the tests
conducted to date have been performed by investigators closely associated with the
aerospace industry. Their temperature range of interest is generally from room tem-
perature to -423 F, with an intermediate test performed at -320 F. Load-compression
strength of a rigid polyester polyurethane foam (2 lb/ft3 ) at these temperatures was re-
ported in Reference 31. At room temperature (77 F) and loads of 25 and 50 psi, the
foam compressed 2.3 and 5.0 percent, respectively. For the same loading conditions
at -320 F, the compression was 15.9 and 26.0 percent. At -423 F and a load of 25 psi,
the compression was 10.9 percent, whereas at a load of 50 psi, the foam partially col-
lapsed and had a compression of 82.5 percent.

A summary of urethane-foam properties at room temperature for various densities


is presented in Table 25 (29) . The compressive strength reported in this table is based
on a specimen deflection of 10 percent. For insulating an LNG pipeline, either the foam-
in-place or the closed-cell rigid urethane foam would be more applicable than the open-
cell flexible foam.

The prices for typical urethane foams depend on density and quantity. Com-
pounded, two-can foam-in-place systems for rigid foam run from 40 to 80 cents per
3
pound. Rigid urethane foam slabs in densities of 1.5 to 2.0 lb/ft will cost from 10 to
19 cents per board foot.

Polystyrene Foams
Polystyrene foam is the most widely used of all the foamed plastics and is avail-
able in a variety of forms - from a basic raw material for molding or extrusion to
already-extruded logs, planks, and boards. A major disadvantage of polystryene is
that it cannot be foamed-in-place.
87

FIGURE 63. VARIATION OF MECHANICAL PROPERTIES WITH


DENSITY OF URETHANE FOAM ( 2 9 )

FIGURE 64. THERMAL CONDUCTIVITY OF URETHANE AS


A FUNCTION OF DENSITY(29)
88

FIGURE 65. EFFECT OF TEMPERATURE ON THERMAL CONDUCTIVITY


FOR VARIOUS INSULATION MATERIAL ( 3 0 )

FIGURE 66. TENSILE PROPERTIES OF URETHANE FOAMS AS


A FUNCTION OF TEMPERATURE ( 3 1 , 3 2 )
90

Three different types of cellular polystyrene are presently available (extruded,


molded, and self-expanded) and although these types differ in technique of manufacture,
most of their properties are quite similar as seen in Table 26(33).
(33)
TABLE 26. PROPERTIES OF POLYSTYRENE FOAMS

Thermal conductivity of molded polystyrene foam is given as a function of density


in Figure 67 ( 2 9 ) . The optimum density from thermal considerations appears to be
2.0 lb/ft 3 . Figure 67 indicates that for this density polystyrene foam has a slightly
higher k-factor than urethane foam.

Figure 68 compares the thermal conductivity of two polystyrene foams as a func-


tion of mean temperature (34) . The fire-retardent foam (designated as Foam A in Fig-
ure 68) is a poorer insulator than the polystyrene foam without these additives (given as
Foam B in Figure 68).

Epoxy Foams

Epoxy foams are characterized by high strength-to-weight ratio, low water ab-
sorption and water-vapor transmission, heat resistance up to 250 F, and self-bonding
qualities to metal surfaces. Because epoxy foams are more expensive than the urethane
foams, they can compete with urethane foams only in those applications where rigid re-
quirements of dimensional stability, uniform density, and moisture resistance are
essential.

The newer liquid-resin foam-in-place formulations have a minimum density of


2 lb/ft3 and a thermal conductivity comparable to that of urethane. Depending on the
manufacturing technique used, epoxy-foam density may vary between 2 and 38 lb/ft3,
and its compressive strength may vary between 15 psi for the lighter density foam and
6000 psi for the heaviest foam presently available. The relationship between com-
pressive strength and density of prefoamed epoxies is given in Figure 69(33). Tensile
strength is also dependent upon density and varies from 30 to 500 psi. Figure 70
91

FIGURE 67. THERMAL CONDUCTIVITY OF MOLDED POLYSTYRENE FOAM


AS A FUNCTION OF DENSITY(29)

FIGURE 68. THERMAL CONDUCTIVITY OF POLYSTYRENE FOAMS ( 3 4 )


92

FIGURE 69. RELATIONSHIP BETWEEN COMPRESSIVE STRENGTH AND


DENSITY OF PREFOAMED EPOXIES ( 3 3 )

FIGURE 70. TENSILE AND SHEAR STRENGTH OF EPOXY FOAM AS


A FUNCTION OF TEMPERATURE ( 3 1 )
93

presents the tensile and shear strengths of an epoxy foam (4.2 lb/ft3) as a function of
temperature (31) . For a temperature decrease from 100 F to -320 F the tensile strength
increases slightly from 52 to 57 psi; for the corresponding temperature range the shear
strength changes from 34 to 32 psi.

Room-temperature physical properties of epoxy foams are presented in


Table 27 ( 3 3 ) . The thermal conductivity at room temperature of the foam-in-place
system is approximately 0.24 Btu-in./(hr)(ft 2 )(F) for the lower densities.
(33)
TABLE 27. PROPERTIES OF EPOXY FOAMS

Silicone Foams

Silicone room-temperature foams are rigid or semirigid with densities from 3 to


5 lb/ft 3 . They are prepared by mixing two silicone resins in the presence of a catalyst.
The reaction is complete in 15 minutes, and maximum strength is developed in 24 hours.

Silicone foams have excellent thermal stability and are either self-extinguishing or
will not burn. Their main applications are in the electrical field and as thermal insula-
tion for temperatures up to 650 F. Silicone foams have high resistance to water absorp-
tion and water-vapor transmission, but they also have low tensile and compressive
strength and may not be suitable for the construction of an underground LNG transfer
line where radial loads may crush the insulation. However, because there is little in-
formation on the mechanical behavior of these foams at temperatures below 77 F, they
cannot be ruled out at the present moment as being unsuitable for proposed application.
94

Typical properties of silicone resin foams (room-temperature-cured) are pre-


sented in Table 28(33). The thermal conductivity at 77 F is 0.281 Btu-in./(hr)(ft 2 )(F),
for a density of 3.5 lb/ft3. The k-factor for silicone foam at room temperature is
higher than the k-factor for either the epoxy or the urethane foam of the same density.

TABLE 28. TYPICAL PROPERTIES OF SILICONE-RESIN FOAMS ( 3 3 )

Phenolic Foams

Phenolic foams of the “reactive” type involve a chemical reaction between phenol,
formaldehyde , and catalyst, and the liberation of a gas by a blowing agent. Since
phenolic foams are a rigid thermoset open-cell product, they absorb water and transmit
water vapor very readily and have good thermal insulating qualities only if the exposed
surfaces are coated.

Densities are reproducible as low as 0.3lb/ft3 but the practical limit is 1.5 lb/ft 3
Strength properties are given as a function of density in Figure 71(33). The phenolic
foams act like brittle solids, failing at strains from 2 to 5 percent.

The thermal conductivity of phenolic foams at various densities for a mean tem-
perature of 100 F are presented in Figure 72 ( 3 3 ) . At an optimum density of 3 lb/ft3,
the thermal conductivity is 0.20 Btu-in. / (hr)(ft 2 )(F). As indicated in Figure 65, the
thermal conductivity of phenolic foam approaches that of urethane foam at a mean tem-
perature of -100 F.

In many cases it is desirable to have a protective hard phenolic skin formed


around a core of more friable rigid foam. By lowering the temperature of the mold
in which foaming takes place, it is possible to form such a skin. However, an ad-
hesive is required when bonding phenolic foam to a smooth hard surface such as found
on glass or metals.
95

FIGURE 71. STRENGTH PROPERTIES OF PHENOLIC FOAM AS


A FUNCTION OF DENSITY(33)

FIGURE 72. RELATIONSHIP BETWEEN THE THERMAL CONDUCTIVITY OF


PHENOLIC FOAMS AND DENSITY(33)
96

Syntactic Foams

Syntactic foams consist of microscopic hollow spheres which are glass, phenolic,
urea, or epoxy mixed with a binding material such as a polyester or an epoxy resin.
The syntactics have high compressive strength and are easy to apply by hand, but can-
not be produced in low densities. These foams according to the proportion of spheres
and binder may have a density ranging from 10 to 68 lb/ft3.

The syntactic foams are useful as a hard facing or protection for rigid foams with
low impact strength. As a class, syntactics have low water absorption and low water-
vapor transmission. The physical characteristics generally approach those of the plas-
tics used as the filler material. Physical properties of syntactic foams are presented
in Table 29 (30) .
(30)
TABLE 29. PHYSICAL PROPERTIES OF SYNTACTIC FOAMS

Cellular Glass

Cellular glass is made by heating pulverized glass with a gasifying substance to a


temperature that causes the glass to flow. As the gas expands it forms hollow cells or
voids which amount to about 90 percent of the material. After the glass has completed
expansion into a closed container, it is removed in the form of rigid blocks and then cut
to the desired shape.

Cellular glass has good thermal efficiency, it is noncombustible, and each cell
provides an effective barrier against convection currents and moisture diffusion.

The thermal and strength values for a cellular-glass density of 9 lb/ft3 are pre-
sented in Table 30(35). The thermal conductivity and strength values listed in Table 30
are room-temperature values. Low-temperature properties could not be found for cel-
lular glass.

If the LNG pipeline is insulated in the field, preformed sections of cellular glass
could be used. This material would require a 7-inch thickness for a heat loss equivalent
to 3 inches of urethane foam (6 lb/ft3 density).
97

TABLE 30. CELLULAR-GLASS INSULATION THERMAL


AND STRENGTH VALUES (35)

Cork and Balsa Wood


Corkboard and balsa wood have been investigated for insulating liquid-hydrogen
missile and space vehicle tanks ( 3 6 ) . Balsa wood has also been used to insulate liquid-
methane tanks.

Cork insulations are made from granules of pure cork which are compressed and
baked. Baking releases natural resins in the material which act as an adhesive and
bind the granules together. Corkboard insulations, intended primarily for low-
temperature applications have a thermal conductivity of about 0.26 Btu-in./(hr)(ft 2 )(F)
for a density of 7.0 lb/ft3. Allowable operating temperatures range from -300 to +200 F.
If the lowest possible conductivity is desired, the insulation must be well sealed from
the influx and condensation of ambient air. Otherwise unsealed corkboard can very
easily exhibit a thermal conductivity higher by at least 50 percent than when it is sealed.
Figure 73 presents the variation of thermal conductivity with temperature for corkboard
having a density of 20 lb/ft 3(37) .

FIGURE 73. THERMAL CONDUCTIVITY OF CORKBOARD AS A FUNCTION


OF TEMPERATURE ( 3 7 )
98

Balsa wood combines low conductivity with comparatively high structural strength.
Table 31 presents the physical properties of this material for two densities (35) . For a
density of 5 lb/ft 3 , the k-factor is 0.29 Btu-in./(hr)(ft 2 )(F) at 60 F. This thermal con-
ductivity is slightly higher than that indicated for corkboard (7.0 lb/ft 3 ). The com-
pression and tensile properties are dependent on the direction of loading and are
presented in Table 31 for loading perpendicular to the grain. Balsa wood, similar to
corkboard, has a high rate of water absorption and must be protected with vapor barrier
in order to retain its thermal-conductivity properties.
(35)
TABLE 31. PHYSICAL PROPERTIES OF BALSA WOOD

Vacuum Insulations

H i g h V a c u u m

Vacuum-type insulations have been used for many years to insulate tanks for the
storage of liquefied oxygen, nitrogen, and hydrogen. However, this type of insulation
has not been used extensively for piping systems due to the necessity of periodic pumping
for the maintenance of a high vacuum. Also, the high cost of vacuum-jacketing piping,
which may run as high as $200 per linear foot for a lo-inch-diameter pipe, may make
this type of insulation uneconomical for the construction of an LNG line. Both the inner
pipe and the outer jacket are usually constructed from Type 304 stainless steel.

In high-vacuum insulation, ideally there is no heat transfer by conduction through


solids, and gas conduction is made negligible by having a low pressure, normally in the
neighborhood of 10 -6 mm mercury. However, in reality, some heat conduction through
solids does occur because of the necessity of using spacers to locate the inner pipe con-
centrically within the outer jacket.

In vacuum-jacketed pipe installations for a line carrying liquid hydrogen, the an-
nular space is usually evacuated to a pressure of 10-2 mm mercury warm. This
relatively “soft” vacuum is sufficient for a liquid-hydrogen line because of the effect of
cryopumping at low temperatures. Cryopumping is effective at liquid-hydrogen tem-
perature (-423 F) because as the inner line cools down the various gases in the vacuum
space except hydrogen are frozen out which greatly increases the vacuum. Since
gaseous hydrogen is less than 0.01 percent of the original air, the pressure due to
cryopumping is reduced to 10 -6 mm mercury and thermal conductivity is approximately
1.4 x 10 - 2 Btu-in./(hr)(ft 2 )(F).
99

At -259 F only carbon dioxide and water vapor would be frozen out of the vacuum
space and the reduction in pressure due to cryopumping would be negligible. If the line
is purged with CO2, the effect of cryopumping would be increased and the resultant cold
vacuum would be somewhere between 10-3 to 10 -4 mm mercury for an initial warm
vacuum of 10 -2 mm mercury. However, at -259 F, cryopumping would not take care of
the long-term permeation through the inner pipe or jacket, nor would it take care of the
released gases from foreign materials within the vacuum space or from the pipe or
jacket metal and periodic pumping would be necessary. For an initial vacuum from
10-3 to 10-4 mm mercury, the thermal conductivity would be somewhere in the vicinity
of 2.7x 10 - 2 Btu-in./(hr)(ft 2 )(F).

The longitudinal contraction due to the temperature difference between the initial
and final operating temperature must be compensated for in a vacuum-jacketed line if it
is to operate successfully. Furthermore, this problem becomes more critical due to
the relative motion between the inner line and the outer jacket when the line is cooled
down to operating temperatures.

Evacuated Powder
Powder insulations consist of finely divided particulate materials such as perlite,
expanded silicone dioxide, calcium silicate, and diatomaceous earth. Phenolic sphere s
are also used. At atmospheric pressure, powder reduces the total heat transport
across the insulation by reducing both convection and radiation. For powder evacuated
to low pressures, gas conduction is negligible and heat transfer is primarily by radia-
tion and solid conduction.

The vacuum required with powder insulations is much less extreme than in the
high-vacuum or multilayer insulations. As a pressure of 10 -2 mm mercury or higher
is adequate, only mechanical vacuum pumps are needed. In order to reduce solid heat
conduction to a tolerable value, these insulations must be relatively thick. There is
also a question of whether the advantage of maintaining a lower vacuum by using evacu-
ated powder instead of high-vacuum insulations offsets the increased fabrication costs
and possible maintenance complications. Because most powders absorb considerable
moisture, they must be thoroughly dried before the required vacuum can be obtained.

If care is not taken in the design and evacuation of the line, several difficulties
may be encountered. Filters (usually installed in the vacuum space) are necessary to
keep the powder out of the pumps. It is necessary to construct the evacuation piping so
that the air will be evacuated at several locations because the powder provides a high
resistance to gas flow. In order to prevent the powder from settling and providing in-
sufficient insulation at the top of the pipe, the powder should be uniformly compacted
prior to completing the fabrication of the lines.

Because powders such as perlite and silicon aerogel are transparent to room-
temperature radiation, metal powders (e. g. , aluminum or copper) are added to the
powder to reflect the radiation. There is an optimum mixture of metal powder and
perlite which will result in the lowest transfer for a specific thickness (37,38) . F i g -
ure 74 shows the typical-performance curve for a mixture of a powder with either cop-
per or aluminum flakes in varying proportions. It is evident from this figure that
opacified powders are four to five times better than conventional powders.
100

FIGURE 74. OPACIFIED POWDER PERFORMANCE ( 3 8 )

FIGURE 75. PERFORMANCE OF EVACUATED POWDER AS


A FUNCTION OF VACUUM ( 3 8 )
101

Figure 75 presents the thermal performance of a powder insulation as a function


(38)
of absolute pressure in the insulating space . Note that a vacuum of 10-2 mm
mercury is quite satisfactory for the evacuated-powder insulation.

Multilayer

Multilayer insulations are composed of highly reflective radiation shields of thin-


gage foil separated by low-conductivity spacer materials such as glass or quartz. In
some cases, multilayers of aluminized polyester film (Mylar) are used without spacers.

There are problems in the use of multilayer composites. Evacuation to relatively


high vacuum is required, necessitating the use of an outer jacket. When evacuated,
these so-called “super-insulations ” have by far the lowest thermal conductivity of any
systems developed to date [10-4 to 10-5 Btu-in./(hr)(ft2)(F)]. Although extremely ef-
ficient from a thermal viewpoint , multilayer insulations suffer from serious physical
and de sign limitations. The structural integrity of two surfaces with an evacuated space
between them is limited by the strength of the supports or spacers. These supports
provide the major heat leak in the system. Also, large evacuated structures are con-
sidered to be inherently unreliable.

Several commercial insulations based on layers of foil and fiber mats or nettings
have been developed. SI-12 consists of 2 to 20 layers per inch of 3-micron glass fiber
mat separated by aluminum foil. This insulation has a density of about 2 lb/ft3 and a
thermal conductivity of 1.1 x 1 0 - 4 Btu-ft/(hr)(ft 2 )(F). Another higher grade insulation
is SI-4. Although it is more costly than SI-12, it has a thermal conductivity of 2.5 x
10 - 5 Btu-ft/(hr)(ft 2 )(F). Figure 76 gives thermal-conductivity values for two multilayer
insulations as a function of absolute pressure in the insulating space. ( 3 8 ) This figure
indicates that an absolute pressure of 10-4 mm mercury is necessary to realize the full
heat-transfer resistance of these multilayer insulations.. At a pressure of 10 -2 mm
m e r c u r y , the SI-12 insulation has a thermal conductivity slightly higher than that in-
dicated for evacuated powder at this pressure (Figure 76).

FIGURE 76, PERFORMANCE OF MULTILAYER INSULATION AS A FUNCTION


OF VACUUM ( 3 8 )
102

VAPOR BARRIER

The primary problem associated with the use of conventional insulations stems
from the requirement of a vapor barrier. Effective low- temperature insulation in-
volves effective moisture control because most insulating materials will absorb water
or transmit water vapor. (This, of course, ceases to be a problem when the temper-
ature of the warmest part of the insulation drops below 32 F. But, since the pipeline
may be operated intermittently it must be considered).

Water vapor which gets into the insulation will condense when cooled to its dew
point. The condensate can solidify and produce a layer with the relatively high thermal
conductivity of the solid. Also, some of the water will freeze inside the insulation and
cause local fissures, thus allowing even more water vapor to enter. These cumulative
effects will soon damage the insulation system within the affected section

The plastic- and glass-foam-type insulations which have a construction with the
majority of the cells closed are impervious and will not absorb water. With these it is
only necessary to seal along the joints with a suitable mastic. However, since not all
the cells will be closed, it is probably a good idea to provide a vapor barrier anyway
Also, moisture that might enter through cracks or unsealed joints will damage the in-
sulation in time. Although the foamed-in-place plastic insulations have no joints re-
quiring sealing, they still should be protected with a vapor barrier, in case cracks de-
velop due to thermal stresses during the initial cool-down period or temperature
cycling during the operation of the line.

The vapor barrier may be any type of material that will retard the absorption of
water vapor by the insulation. Metals, foils, plastics, and various types of coating
may be used, although the most practical and most widely used vapor barriers are a
combination of mastics and membranes.

Tensile and elongation properties of a polyester film, a polyvinyl fluoride film,


and a fluorohalocarbon film were measured at -320 F and are presented in Table 32. (40)
The polyester film appears to be the strongest of the materials tested (Table 32) and
was found to be nondirectional in its properties. Reference 31 reports work carried
out on a 3-mil Mylar, a laminate made of one layer of 1-mil aluminized Mylar, one
layer of 116 glass cloth, and Epon 820 binder (Table 33). The linear coefficients of
expansion of three vapor-barrier materials are shown in Figure 77. (3 1 , 41) The fiber-
glass laminate, because of the stabilizing influence of the inorganic fibers, has the
lowest average expansion of the materials studied.

Table 34 shows the results of thin-film tensile tests. (42) In general, the films
have higher ultimate-stress points at liquid-nitrogen temperatures than they do at room
temperatures. Although the laminate vapor-barrier materials under consideration have
a much higher ultimate-stress point than the film materials, the stress at which they
fail as vapor barriers due to crazing and flaking off of the epoxy resin is from 200 to
400 percent lower than the ultimate-stress points for the film materials Without
losing their properties as vapor barriers, the films have far superior mechanical pro-
perties with regard to stress and strain loads than can be withstood by the laminate
materials
103

TABLE 32. TENSILE AND ELONGATION PROPERTIES OF


VAPOR-BARRIER MATERIALS AT -320 F ( 4 0 )

TABLE 33. TENSILE PROPERTIES OF COMPOSITE LAMINATES AT


ROOM AND CRYOGENIC TEMPERATURES ( 3 1 )
104

FIGURE 77. LINEAR EXPANSION AS A FUNCTION OF TEMPERATURE


FOR THREE VAPOR-BARRIER MATERIALS ( 3 1 , 4 1 )

Extensive investigations have been carried out in order to evaluate the bonding
properties of adhesives at cryogenic temperatures. These could be applied as a mastic
over the outside surface of the insulation to provide a vapor barrier, or they could be
used in conjunction with one of the films previously mentioned. However, the majority
of the adhesives investigated require curing at a temperature around 300 F, and would
not be generally applicable for an LNG line. Some work has been done on room-curing
adhesives and the results appear promising for this particular application.

A new room-temperature-curing polyurethane adhesive has excellent properties


at low temperatures. This two-part system provides high ultimate-shear and tensile
strengths at temperatures as low as -423 F. Figure 78 indicates that this adhesive in-
(43)
creases in strength as the temperature decreases.

An epoxy-polyamide is readily mixed, easily applied, has good potential life, and
can be cured at room temperature to yield a flexible system. Its low- temperature per -
formance, however, is not as good as that of the polyurethane adhesive.
105

FIGURE 78. EFFECT OF TEMPERATURE ON SHEAR STRENGTH OF


A POLYURETHANE ADHESIVE ( 4 3 )

An epoxy-nylon polyamide is another room-curing-temperature adhesive which


has excellent peel properties and shock resistance at cryogenic temperatures. This
two-part paste system can be used over a temperature range from -423 to 160 F

Depending on the type of insulation used, one of the following four methods should
be satisfactory for keeping ground water from entering the insulation and the expansion
chamber if expansion joints or loops are used:

(1) Hot asphalt or tar enamel applied directly over a closed-cell type
of insulation such as urethane foam.

(2) A vapor barrier which consists of alternate layers of hot asphalt


or tar and a 10 by 10 mesh glass cloth applied directly over the
insulation.

(3) A thin-gage aluminum sheath to cover the insulation. Because


aluminum requires a protective waterproof coating, the sheath
can be spot or tack welded and the joints in the sheath filled in
and covered with a mastic prior to coating the line.

(4) A combination of membranes and mastics, such as fiberglass


laminates or thin films with one of the adhesives discussed above.
The procedure involved would necessitate the application of the
adhesive to the insulation followed by wrapping the membrane
around the pipe. It probably would be advantageous to coat this
assembly with either hot asphalt or tar.

In order to provide the best vapor barrier at the lowest cost, additional investiga-
tions which would include some experimental tests are necessary.
106

MISCELLANEOUS EQUIPMENT

Cryogenic Pumping Equipment

The fundamental concepts and methods which are adequate for the fluid-mechanical
design of pumps for common fluids are applicable to liquefied gases The performance
characteristics will differ somewhat because of such things as greater leakage due to
lower viscosity The efficiency may be lower due to the fact that a less-dense fluid re-
quires less total power input while the so-called fixed losses (such as those occuring in
bearings) will not decrease appreciably.

For specified flow rates, Figure 79 indicates the maximum discharge pressures
which may be realized from currently available pumping equipment (44) . Flows from
200 to 500 MMCFD would require liquid flow rates approximately from 1700 to 4200
gallons per minute and thus necessitate the use of centrifugal pumps. This is because
positive-displacement pumps are generally used for low-capacity high-pressure appli-
cations. Most of the design information such as performance curves is considered to be
proprietary information and released only to potential customers. Several manufac-
turers indicated that for the flows and pressures considered in this report, centrifugal
pumps for methane could be built with technical knowledge and capabilities presently
available. Although specific efficiencies were not given, the general consensus of o-
pinion is that the liquid-methane pumps would have efficiencies in the neighborhood of
75 percent. This would need to be confirmed.

FIGURE 79. PRESSURE AND FLOW-RATE CAPABILITIES OF


COMMERCIALLY AVAILABLE CRYOGENIC
PUMPS ( 4 4 )
107

Valves

The problems arising in connection with valves for liquefied gases stem from heat
leak, distortion due to temperature gradients, icing together of arts which must have
(45)
relative motion, freezing of packing, and seat tightness. Beard discusses several
types of cryogenic valves which include shutoff, ball, butterfly, solenoid, gate, and
safety. The article also contains a valve material compatibility guide,

The performance of a cryogenic butterfly valve incorporating a spherical sealing


surface is discussed in Reference 46. Although most butterfly valves leaked severly at
cryogenic temperatures several years ago, butterfly valves are produced today with a
guaranteed maximum leakage of 2.5 seem of helium per inch of diameter at -320 F.

Instrumentation

Instrumentation for cryogenic applications is not significantly different from con-


ventional instrumentation except that the low-temperature, and possibly high-vacuum,
requirements may affect the details of some of the sensing elements. Temperature,
pressure, and flow measurements will be briefly discussed as they are related to an
LNG transfer line.

Temperature

Cryogenic temperatures can be measured by thermocouples, vapor-pressure


bulbs, gas thermometers, resistance thermometers, and semi-conductor thermom-
eters. For a comprehensive coverage of the field of thermometry, the reader is re-
ferred to References 47 and 48.

Thermocouples have been used widely in cryogenic applications. Their main as-
sets are adequate accuracy, good response, and moderate cost. Copper-constantan
junctions have been used satisfactorily from room temperature down to -423 F.

Vapor-pressure thermometers are extremely accurate and simple to operate, but


have limited ranges

Gas-filled thermometers are difficult to use and are undesirable for application in
an LNG piping system.

Resistance thermometers are accurate, stable, and linear down to -423 F, but
they are very expensive.

Semiconductor thermometers depend on the rapid change of resistance with tem-


perature of the metallic-oxide sensing elements for their accuracy. These thermom-
eters are also relatively simple to operate, but have the disadvantages of nonlinear
calibration curves which are nearly exponential. In certain instances, the semicon-
ductor thermometers have been found to be unstable.
Pressure

Line pressure is usually measured through a tap in the pipe wall. All contami-
nants which can freeze and block the pressure taps and connecting tubing must be re-
moved prior to cooling the pipeline. There are pressure transducers available which
are made and calibrated for use down to -320 F and lower. However , if the trans-
ducers are at or near the temperature of the liquefied gases, compensations for the
temperature effects due to the zero shift must be made, which requires a knowledge of
the temperature of the transducer.

If vacuum-jacketed piping is used, a device is necessary to monitor the pressure


within the annular space. This is usually a commercial vacuum gage.

Flow

Turbine meters, orifice meters, and venturi meters have been used to measure
flow. The last two meters require a knowledge of the fluid density in order to obtain
flow measurements because these meters utilize the Bernoulli effect (dependent on pres-
sure-difference reading). Volumetric flow measurements from turbine-type flow
meters can be obtained from the meter alone.

Several two-phase mass-flow meters are presently available, but additional work
is necessary in this area to develop a simple, rugged device for measuring the flow of
a liquid-vapor mixture However , because an LNG pipeline will presumably involve
single-phase liquid methane, the volume measurements from a turbine-type flowmeter
and a knowledge of the density can be used to calculate the mass flow rate.

Care must be exercised during the cool-down of the pipeline in order that the
large volumes of gas generated by flashing of the liquid do not overspeed and damage the
flow meter. The turbine-type flow meters will operate reliably in an LNG pipeline
provided they are not abused.
109

DISCUSSION

Based on the preliminary heat-transfer analysis in this report, indications are


that LNG can be transmitted over fairly long distances. Calculations show that dis-
tances of 25 to 130 miles without intermediate pumping stations and of 100 to 1950 miles
between cooling stations appear possible. Factors such as pipe diameter and flow rate
influence the pumping- and cooling-station spacing. Furthermore, the distance between
cooling stations is very much dependent on the type and thickness of insulation used.

For short transfers of LNG, 10 miles or less, no intermediate pumping or cooling


stations are required. Because the pressure drop and heat influx are fairly low, a
lo-inch-diameter pipe insulated with 2 inches of urethane and a throughput of 200
MMCFD would require a discharge pressure of only 135 psig to prevent flashing The
discharge-pressure calculations are based on a pressure drop of 120 psig and a pres-
sure requirement of 15 psig to insure that LNG does not flash at the terminal point of
a lo-mile transfer line where the LNG temperature is calculated to be -249 F.

The choice of materials is somewhat affected by the transmission distance in that


the pipe wall thickness , based on a design stress of 52,650 psi (72 percent of SMYS) for
9 percent nickel steel in the quenched and tempered condition, would be about 0.013
inch (lo-inch-diameter pipe, 135 psig pressure), which is considerably below the least
nominal wall thickness of 0.104 inch specified in USASI B31.8 for this pipe size.
Therefore, the better choice for short transfer distances and low pressures would
probably be low-yield-strength aluminum alloys such as 3003-0 or 6061-T6.

The cost of pipe material and bellows for a lo-inch line built from 9 percent
nickel steel with a 0 104-inch wall thickness is estimated at approximately $74,600 per
mile. This figure includes the pipe-material cost ($58,000) and the cost of bellows
($16,600)) but excludes insulation and construction costs . For aluminum alloy 6061-T6
the estimated cost is $42,000 per mile, where the material cost is taken as $11,000 per
mile and the cost of bellows is taken as $31,000 per mile. Material costs used in these
calculations are high because prices were available only for orders of 20 tons or less.
Although more expansion loops or bellows are needed for aluminum alloy 6061-T6 than
for 9 percent nickel steel, the saving in material cost compensates for this additional
expense. A possible alternative, of course, is to use 9 percent nickel steel, smaller
pipe sizes, and higher line pressures to transmit the same flow.

It appears from the calculations that for LNG transmission distances of 10 miles
or less, 2 inches of urethane foam on either a lo-inch or 20-inch-diameter pipe would
be sufficient to prevent LNG from flashing anywhere along the line

Results obtained from the heat-transfer analysis indicate that for the insulations
considered, the pressure-drop curve and not the temperature curve is the governing
criterion, and therefore more pumping stations would be required than cooling stations.
This is most fortunate because for long transmission distances there is considerable
economic advantage in keeping the line length below that which would require an inter-
mediate cooling station. It is estimated that an intermediate reliquefaction plant would
have an efficiency in the order of 10 percent and that the cooling horsepower require-
ment would probably be around 90 percent of the total mechanical power necessary for
pumping and cooling.
110

Recall that for Case A (200 MMCFD, lo-inch-diameter pipe, 2 inches of urethane
insulation) the pumping- and cooling-station spacings were 25 and 100 miles, respec-
tively. The transmission distance before the first cooling station is required can be in-
creased from 100 to 125 miles by allowing the LNG temperature to increase to -120 F
instead of - 140 F (used as the limiting temperature in the calculations). However , since
-120 F is near the flashing point of LNG (-116 F), fairly close control of the insulation
and soil thermal conductivities would be necessary to predict at what point along the line
the LNG will reach a temperature of -120 F. Although the temperatures used above are
for liquid methane, and may be slightly different for liquefied natural gas, the same
reasoning still holds. Additional insulation would permit greater cooling-station
spacing, but there is a break-over point beyond which it might be less expensive to in-
stall and operate a cooling station than to increase the insulation thickness further.

For short transfer distances and line pressures less than 250 psig, either bel-
lows or expansion loops can be used to compensate for the longitudinal thermal contrac-
tion of the line. A line constructed from 9 percent nickel steel would require 37 bellows
per mile, where the axial movement of each bellow is computed to be 2.94 inches. The
number of bellows required per mile of pipeline increases to 53 for Type 304 stainless
steel and to 69 for aluminum.

Bellows are not recommended for an LNG line operating at 1000 psi because bel-
lows subjected to this pressure become expensive, have limited axial movement (1 inch
or less for either a 10-inch or 20-inch line) and their long-term reliability at this pres-
sure cannot be accurately ascertained. Therefore, either expansion loops or a re-
strained Invar line must be used to compensate for line contraction. Because of the
present-day high cost of Invar (almost twice as much as for 9 percent nickel steel), it
appears that expansion loops would be better for a long-distance LNG transmission
line.

A large number of possible candidate pipe materials have been examined in this
report. In selecting the most promising material for the construction of an LNG line
the following important factors were considered: material design stress, cost of the
material, and the material thermal-expansion coefficient. The three materials which
will be discussed presently appear to have suitable fracture-toughness characteristics;
however, additional testing is needed for verification of these properties.

The materials which appear to be the best candidates are 9 percent nickel steel,
Type 304 stainless steel, and 5083-0 aluminum. Although there are other aluminum
alloys which could be used for the construction of an LNG line, 5083-0 aluminum is
taken as the representative material for this alloy group. Invar, which has some design
advantages, will not be considered here because of its high cost.

If the design stress is taken as 72 percent of SMYS, then 9 percent nickel steel in
the quenched and tempered condition has a design stress of 54,000 (0.72 x 75,000) psi.
The design stress for Type 304 stainless steel is 21,600 (0.72 x 30,000) psi and that for
5083-0 aluminum is 12,240 (0.72 x 17,000) psi. Considering the design stress of Type
304 stainless steel as unity, then the design stress ratios for 9 percent nickel steel and
for 5083-0 aluminum are 2.50 and 0.57, respectively.
111

For a design utilizing expansion loops, additional material over that necessary
for a straight run of pipe is required in order to construct the legs of the loops. The
results from Table 3 indicate that this additional material amounts to 16.0 percent of
the material necessary for a line constructed from Type 304 stainless steel without ex-
pansion loops. The corresponding figures for 9 percent nickel steel and 5083-0 alu-
minum are 6.6 percent and 14.8 percent, respectively.

Material costs for large-quantity orders could not be obtained. The cost data
given is for 10-inch-diameter pipe purchased in quantities of 20 tons or less. Type 304
stainless steel, 9 percent nickel, and 5083-0 aluminum cost $0.75/lb, $0.91/lb, and
$0.77/lb, respectively, These cost figures are presented for illustrative purposes only,
because the cost of the material is somewhat dependent on the wall thickness specified.

The design stress, additional material used for expansion loops, and material
costs can be incorporated into a single ratio. If the cost of an LNG line constructed
from Type 304 stainless steel is taken as 100 percent, then the cost of a line constructed
from 9 percent nickel steel and 5083-0 aluminum would be 45.4 percent and 61.3 per-
cent, respectively. Because the design and construction of an LNG line concerns not
only the relative strengths and cost of the materials, but ease and reliability of fab-
rication, it is felt that the cost comparison made above should be used with caution. Of
the three materials discussed, 9 percent nickel steel in the quenched and tempered con-
dition appears to be the best candidate material from economic considerations for the
construction of a long LNG transfer line.

There are possibilities of using lower cost materials such as 3-1/2 percent
nickel steel, but additional testing is necessary to establish the suitability of the
fracture-toughness properties of these materials for an LNG line. A line constructed
from 3-1/2 percent nickel steel would cost 43.5 percent as much as a line constructed
from Type 304 stainless steel. The relative cost figure of 43.5 percent is based on de-
sign stress of 28.800 (0.72 x 40,000) psi, a requirement of 12.1 percent of additional
material for expansion-loop construction, and a material cost of $0.45/lb.

The relative{-cost ratio for 3-1/2 percent nickel steel is only slightly lower than
that for 9 percent nickel (43.5 vs 45.4 percent), This stems from the fact that for a de-
sign stress based on 72 percent of SMYS, 9 percent nickel in the quenched and tempered
condition has a design stress 1.88 times that of 3-1/2 percent nickel. In this case the
higher design stress of the 9 percent nickel tends to offset the higher costs of this
material. However, if the design stress is not based on 72 percent of SMYS, but is
based, for example, on USASI B31.3, the relative cost figure for 3-1/2 percent nickel
would be considerably below that for 9 percent nickel because the design stress ratio
decreases from 1.88 to 1.38, but the material costs remain the same.

An analysis of various insulation systems and materials has indicated that although
vacuum-jacketed insulations have thermal conductivities which may be more than 100
times lower than the conductivities of mechanical-type insulations, the cost of a
vacuum- jacketed line may be as much as 10 times the cost of a line using conventional
insulation. Due to the high installation cost and the probable requirement for periodic
pumping to maintain the vacuum, vacuum-jacketed insulations are not recommended for
an underground LNG pipeline.
112

Of the conventional insulations discussed, urethane-foam insulation appears to


have the best combination of thermal-performance characteristics, strength, ease of
fabrication, and low cost. For an equivalent thermal resistance to heat transfer as
provided by 2 inches of urethane foam on a lo-inch-diameter pipe, it would require
2.40 inches of phenolic foam ($0.25), 2.83 inches of polystyrene foam ($0.12), 3.25 in-
ches of cork ($0. 17), 3.75 inches of epoxy foam ($0.30), 6.25 inches of balsa wood
($0. 24), and 6.25 inches of cellular glass ($0. 15). The figures in brackets represent
the cost of the material in dollars per board foot at approximately optimum densities
from heat-transfer considerations. The cost of urethane foam for a density of 2 lb/ft3
was obtained as $0.18/bd ft.

If the cost of 2 inches of urethane on a lo-inch pipe is taken as unity, then the
other insulating materials exhibiting equivalent thermal performance would have cost
ratios as follows: polystyrene foam (1.01), cork (1.68), phenolic foam (1.71), cellular
glass (3.79), epoxy foam (4.53), and balsa wood (6. 06). It appears from these cost
ratios that urethane foam and polystyrene foam are the least expensive insulations on the
basis of material cost and material requirements. Installation costs are somewhat de-
pendent on the thickness of insulation used and therefore urethane foam is preferred
over polystyrene foam because, for equivalent heat-transfer resistance, polystyrene
foam would require a thicker insulation wall (2.8 inches versus 2.0 inches on a lo-inch-
diameter pipe).

If urethane foam is used as the insulating material, consideration must be given


to the most economical method of insulating the pipe. For a density of 2 lb/ft 3 , foam-
in-place urethane costs somewhere in the vicinity of $0.80/lb, whereas urethane billets
or boards may cost approximately $1.10/lb

For a line insulated in the field., preformed sections of urethane could be used.
However, these sections would probably be fabricated from urethane billets or boards
and fabrication costs in addition to the cost of the material would be involved. The
vapor-barrier material and installation costs must be added on to the insulation costs.

Joints of pipe could be shop-insulated with foam-in-place insulation, covered with


a vapor barrier and then shipped to the construction site, thus reducing field-
construction time. One firm has indicated that for shop insulating a lo-inch-diameter
pipe with 2 inches of urethane foam, the cost would be $7.35 per foot of pipe. This
price includes an outer jacket (110 mil extruded rigid vinyl) which is bonded to the in-
sulation and acts as a vapor barrier.
113

RECOMMENDATIONS FOR FUTURE WORK

In carrying out the work required for this report, some future steps in the re-
search and development of an LNG pipeline have become evident. Accordingly, future
work in the following areas is recommended:

(1) Additional work on design


(2) Laboratory determinations of mechanical and thermal properties
of several material candidates
(3) Full- scale fracture tests to evaluate resistance to fracture
initiation and propagation on promising pipe materials
(4) Construction and testing of an “Athens” loop,

Additional Work on Design

The major emphasis of the work being reported here has been on materials of
construction. Recognizing that material considerations could not be entirely divorced
from design considerations, enough work was carried out relative to design to establish
a few basic configurations, Although it is not intended to produce a single final design
of an LNG pipeline since several feasible alternatives seem likely, a few areas are
considered universally important and worthy of additional work.

One of these involves the early transient period during which the warm line is
being initially cooled. The time for cool-down and filling of cryogenic piping in past
applications has been successfully predicted by static- equilibrium methods which con-
sider only end-point conditions. However, the rate of heat transfer and the internal
fluid dynamics have not previously been considered. Thus, these simplified methods
fail to tell anything about the dynamics of filling and such parameters as possible over-
pressure or required local venting, Of course, the longer and bigger the line, the
more important these dynamic effects become. Additionally, because of the low vis-
cosity of LNG, liquid may advance well ahead of the location where the pipe is flowing
full. This may be a problem in hilly country where the combined effect of liquid “run-
ahead” and heat-influx may produce “ s l u g s " of liquid which could, under certain con-
ditions, attain high velocities.

Another design problem which appears to warrant further consideration is that of


frost-heave. Frost-heaving results when a clay-like or silty soil “grows” ice lenses
in one or more layers. Moisture content, degree of surface tension around soil grains,
permeability, and amount of capillary potential are factors affecting susceptibility to
frost-heave. Ice lenses develop with their long dimensions perpendicular to the di-
rection of temperature gradients and, therefore, would develop in a frost-heave-
susceptible soil parallel to the pipe and would exert radial pressures on the pipe. A d -
ditional work is recommended to (1) determine whether these potential pressures can
be deleterious and (2) determine the most appropriate methods of alleviating these
pressures.
114

Other design areas which warrant further investigation are (1) an analysis of
bowing stresses incurred from temperature gradients when an LNG pipeline flows
partially full and (2) some additional design work on hold-down anchors.

Laboratory Determination of Mechanical and Thermal

Properties of Several Material Candidates

Material suppliers are becoming increasingly aware of the need to know the low-
temperature properties of their materials. However, many of the property data which
are considered important in the design of a cryogenic pipeline are either not available
or are of a type which are not directly applicable. Examples of these deficiencies for
pipe materials are (1) the lack of thermal-contraction coefficients, (2) the lack of ap-
propriate fracture-toughness data, e. g. , sometimes only Charpy keyhole-notch energy
levels at various temperatures are given, and (3) the lack of strength properties at the
expected service temperature (-259 F in this case).

Some similar deficiencies have been observed for insulating materials. Data for
insulating materials are, however , generally much more inclusive.

The work done this year has narrowed the choice of acceptable materials through
a paper study of available data. It is recommended that for some of the promising
candidate materials for which either the important data is unavailable, or, if available,
requires confirmation, mechanical and thermal properties be experimentally deter-
mined. For a few pipe materials, it is suggested that appropriate fracture-toughness
data, strength- at- temperature data, and contraction- coefficient data be obtained.
Similarly, for insulation materials, it appears advisable to obtain some conductivity
data (over a temperature range), some contraction data (over a temperature range), and
some strength data including elastic constants.

Full-Scale Fracture Tests to Evaluate Resistance

to Fracture Initiation and Propagation

As pointed out in the section, “Resistance to Fracture”, (see page 39), fracture-
toughness properties are important from two viewpoints: propagation resistance and
initiation resistance. From the propagation viewpoint, as was discussed, the relation
between fracture speed and the speed of decompression are of prime importance. A l -
though laboratory impact tests are instructive and useful in estimating fracture
characteristics, it is recommended that the actual fracture characteristics be deter-
mined by full-scale tests.

Also, it is apparent from this year’s work that there is insufficient valid informa-
tion about fracture-initiation resistance of the important pipe materials.

It is because of this reasoning that some full-scale West Jefferson-type tests are
recommended for the important material candidates, particularly those which are not
as costly but which may have transition temperatures higher than the anticipated service
temperature. Briefly, the West Jefferson-type tests consist of welding end caps on a
section of pipe, introducing an artificial defect, cooling to test temperature, and
115

pressuring to failure. Timing wires are placed along the pipe to measure fracture
speed. These tests are capable of evaluating fracture toughness from both initiation
and propagation viewpoints.

Athens Loop

It is recommended that, when the above research is accomplished, the construc-


tion and testing of an experimental loop be undertaken. Although such a loop is re-
ferred to as an “Athens” loop, it couId be constructed elsewhere and might preferably
be located closer to an LNG source.

The proposed LNG pipeline loop would offer the opportunity to fully develop the
concept of transmitting LNG through pipelines before actual commercial projects are
undertaken, Briefly, the following investigative areas are considered relevant to the
proposed loop:

(1) Construction and assembly techniques, Buried cryogenic pipelines


do not have precedence. A facility providing an opportunity to con-
struct an LNG pipeline using various design and construction in-
novations appears de sir able.

(2) Flow characteristics, The loop facility would provide an “actual


service environment” in which flow studies under various condi-
tions of temperature and pressure would be made. Actual mea-
surements of LNG properties (e. g. , viscosity) would also be made.
(3) Heat transfer. For various insulating schemes and flow rates, the
heat-transfer characteristics would be investigated.
(4) Frost-heave. This phenomenon would be directly investigated and
methods of circumventing it would be developed and tested.
(5) Fracture characteristics. Confirmation of fracture characteristics
under actual service conditions would be made, Particularly rele-
vant would be the measurement of the LNG decompression velocity
and the fracture velocity,
(6) Equipment testing, Equipment, such as valves and pumps, would
be studied under service environment.
116

REFERENCES

(1) Jacobs, R. B. , “Single-Phase Transfer of Liquefied Gases”, National Bureau


of Standards Circular 596 (December 1, 1958).

(2) Haettinger, G. C. , “Considerations in the Design, Selection and Use of Vacuum


Insulated Pipe”, Source of article not determined (August 9, 1965).

(3) “Foamglas Application Specifications for Underground Piping”, Pittsburgh-


Corning Corporation.

(4) “Line Expansion and Flexibility”, Tube- Turns Division of Chemtron Corporation.

(5) “Bellows Design Manual”, Tube-Turns Division of Chemtron Corporation.

(6) McClure, G. M., Duffy, A. R., and Eiber, R. J., “Fracture Resistance in Line
Pipe”, p resented at the Petroleum Mechanical Engineering Conference, Los
Angeles, California, September 20-23, 1964, of The Americal Society of Mechani-
cal Engineers and published in Journal of Engineering for Industry, ASME Trans-
actions (August, 1965).

(7) Eiber, R. J., “Correlation of Full-Scale Tests with Laboratory Tests”, Sym-
posium on Line Pipe Research (a progress report on Project NG-18 being carried
out by the Pipeline Research Committee at Battelle Memorial Institute), Pipeline
Research Council International, Inc. publication, Catalogue No. L 30,000, 83-118
(November 17-18 1965).

(8) Van Itterbeek, A., Verbeke, O., and Staes, K., “Measurements on the Equation
of State of Liquid Argon and Methane up to 300 kg cm-2 at Low Temperatures”,
Physica 29, 742-754 (1963).

(9) Duffy, A. R., “Studies of Hydrostatic Test Levels and Defect Behavior”, Sympo-
sium on Line Pipe Research (progress report on Project NG-18 being carried out
by the Pipeline Research Committee at Battelle Memorial Institute), Pipeline Research
Council International, Inc. publication, Catalogue No. L 30,000, 139-160 (November
17-18 1965).

(10) “Mechanical and Physical Properties of the Austenitic Chromium-Nickel Stainless


Steels at Subzero Temperatures”, The International Nickel Company.

(11) “Low Temperature and Cryogenic Steels, Materials Manual”, United States Steel.

(12) Kaufman, J. G., and Wanderer, E. T., “Aluminum for Cryogenic Applications"’,
Machine Design, 199-203 (November 11, 1965).

(13) “Effects of Low Temperatures on Structural Materials”, NASA SP-5012 Tech-


nology Utilization Report, National Aeronautics and Space Administration,
Washington, D. C. (December, 1964).

(14) Bell, J. H. , Jr. , Cryogenic Engineering, Prentice-Hall, Inc., Englewood Cliffs,


New Jersey (1963).
117

(15) “Cryogenic Applications of Alcoa Aluminum”, Aluminum Company of America.

(16) National Bureau of Standards Cryogenic Engineering Laboratory (unpublished


data).

(17) Bogardus, K. O., Stickley, G. W., and Howell, F, M., “A Review of Informa-
tion on the Mechanical Properties of Aluminum Alloys at Low Temperatures”,
N, A. C. A. Technical Note 2082 (May, 1950).

(18) Zambrow, J. L., and Fontana, M. G., “Mechanical Properties, Including Fa-
tigue of Aircraft Alloys at Very Low Temperatures”. Trans. Am. Soc., Met.,
V 41, 480 (1949).

(19) Advances in Cryogenic Engineering, edited by K. D. Timmerhaus, Vol. 9,


Plenum Press Inc. , New York, New York (1964), “Performance of a New Cryo-
genie Aluminum Alloy 7039” (F. W. DeMoney), pp 112-123.

(20) Gatsek, L. F., “Joining Aluminum to Stainless Steel”, Welding Design and
Fabrication, 61-62 (September, 1965).

(21) McClintock, R. M., and Gibbons, H. P., “Mechanical Properties of Structural


Materials at Low Temperatures, National Bureau of Standards, Monograph 13
(June 1, 1960).

(22) Mounce, W. S., Crossett, J. W. , and Armstrong, T. N. , “Steels for the Con-
tainment of Liquefied Gas Cargoes”, Trans. of the Society of Naval Architects,
Vol. 67, 1959, New York, New York.

(23) Armstrong, T. N., Gross, J. H. , and Brien, R. E., “Properties Affecting


Suitability of 9 percent Nickel Steel for Low-Temperature Service”, Welding
Journal Research Supplement, 57 s (February, 1959).

(24) “Steels for Low Temperature Applications”, United States Steel (September, 1961).

(25) Z i c k , L . P . , Crossett, J. W., and Lankford, Jr., W. T., “Destructive Tests


of 9 Percent Nickel-Steel Vessels at -320 F”, ASME paper number, 62-WA-273
(1962).

(26) Gottlieb, T. , and Shira, C. S. , “Fabrication of Iron-Nickel Alloys for Cryogenic


Piping Service”, Welding Research Supplement, 116-S to 123-S (March, 1965).

(27) Schwartzberg, F. R., Osgood, S. H. , Keys, R. D., and Kiefer, T. F., “Cryo-
genie Materials Data Handbook”, Report ML-TDR-64-280, The Martin Company,
Denver, Colorado (January, 1964).

(28) Rolfe, S. T., Haak, R. P., and Gross, J. H., “Structural Suitability of a High-
Toughness Alloy Plate Weldment with a Minimum Yield Strength of 140 ksi”,
Welding Research Supplement, 40s-48s (January, 1965).

(29) Modern Plastics Encyclopedia, McGraw-Hill, New York, New York, 375
(September, 1965).
118

(30) Gerstin, H., “How to Evaluate the Rigid Plastic Foams”, Product Engineering,
59-68 (June 21, 1965).

(31) Advances in Cryogenic Engineering, Edited by K. D. Timmerhaus, Vol. 8,


Plenum Press, Inc., New York, New York (1963) (R. N. Miller, et al, ), pp
417- 424.

(32) Campbell, M. D., Haskins, J. R., Hertz, J., Jones, H, , and Percy, J. L.,
“Thirty Day Evaluation of Foams and Honeycomb for Centaur Intermediate
Bulkhead”, MRG-312 (April 26, 1962). Contract AF 33(616)-7984 (Compilation
of Materials Research Data). (AD 291 520). CEL-NBS 16160.

(33) Plastics Engineering Handbook, Society of the Plastics Industry, Inc. , 3rd Edi-
tion, Reinhold Publishing Corporation, New York, New York (1960).

(34) Advances in Cryogenic Engineering, Edited by K. D. Timmerhaus, Vol. 7,


Plenum Press, Inc., New York, New York (1962) (J. F. Haskins and J. Hertz),
pp 353-359.

(35) Fabian, R. J., “Thermal Insulation Materials”, Materials in Design Engineering


(March, 1958).

(36) Gray, V. H. , Gelder, T. F., Cochran, R. P., and Goodykoontz, J. H., “Bonded
and Sealed External Insulations for Liquid-Hydrogen-Fueled Rocket Tanks during
Atmospheric Flight”, NASA TN D476, Washington, D. C. (1960).

(37) Technology and Uses of Liquid Hydrogen. Edited by R. B. Scott, W. H. Denton,


and C. M. Nicholls, Macmillan Co. , New York, New York (1964), “Thermal
Insulation, Storage, Transport, and Transfer of Liquid Hydrogen” (R. B.
Jacobs), pp 106-148.

(38) Advances in Cryogenic Engineering, Edited by K. D. Timmerhaus, Vol. 5,


Plenum Press, Inc., New York, New York ( 1960) (M. M. Fulk), p 146.

(39) “Low Temperature Technology - What’s the Latest”, Chemical Engineering,


77-82 (November 30, 1959).

(40) Miller, R. N., Bailey, C. D., Freeman, S. M., Beall, R. T., and Coxe, E. F.,
“Properties of Foams, Adhesives and Plastic Films at Cryogenic Temperatures”,
Industrial and Engineering Chemistry, Product Research and Development, 1 (4),
257 (December, 1962).

(41) Bailey, C. D. , “Linear Thermal Expansion of Cryogenic Materials at Cryogenic


Temperatures”, ER-5682 (May 21, 1962) - GEL-NBS 14801.

(42) Advances in Cryogenic Engineering, Edited by K. D. Timmerhaus, Vol. 10,


Plenum Press, Inc., New York, New York (1965), “Low Temperature Tensile,
Thermal Contraction, and Gaseous Hydrogen Permeability Data on Hydrogen
Vapor Barrier Materials” (R. P. Caren, R. M. Coston, A. M. C. Holmes, and
F. Dubus).
119 and 120

(43) Kuno, J. K. , “Adhesive Bonding of Insulation for Temperature Extremes -


Cryogenic to Reentry”, Proceedings 7th National Symposium on Adhesives and
Elastomers for Environmental Extremes, Section 11, 36 (May, 1964).

(44) Carter, T. A. , Jr., and Brown, W. H. , “Cryogenic Pumping Equipment”,


Cryogenic Technology, January-February, 67-70 (1965).

(45) Beard, C. S. , “Control Valves for Cryogenic Fluids”, Control Engineering, 13


(3), 67-72 (March, 1966).

(46) Hoof, R. G. , “Cryogenic Butterfly Valves”, Cryogenic Technology, 160-162


(May-June), 1965.

(47) Scott, R. B. , Cryogenic Engineering, D. Van Hostrand Co. , Inc. , Princeton,


New Jersey (1959).

(48) Advances in Cryogenic Engineering, Edited by K. D. Timmerhaus, Vol. 8,


Plenum Press, Inc., New York, New York (1963), “Temperature Measurements
in Cryogenic Engineering” (R. J. Corruccini), pp 315-333.

(49) McAdams, W. H. , “Heat Transmission”, Third Edition, McGraw-Hill Book


Company, Inc. , New York, New York (1954), p 25.
This Page Intentionally Left Blank
APPENDIX
MATHEMATICAL DERIVATION FOR TEMPERATURE
PROFILE ALONG AN LNG PIPELINE
This Page Intentionally Left Blank
A-1

APPENDIX

MATHEMATICAL DERIVATION FOR TEMPERATURE

PROFILE ALONG AN LNG PIPELINE

Nomenclature

= heat transferred through pipe wall, Btu/lb

a = subscript denoting initial state of interest

b = subscript denoting final state of interest

p = absolute pressure, lb/ft 2

v = specific volume, ft3 /lb

J = 778 ft-lb/Btu

u = internal energy, Btu/lb

V = fluid velocity, ft/sec

g = gravitational acceleration, ft/sec 2

Z = elevation above datum, ft

= net mechanical work, Btu/lb

c = specific heat, Btu/(lb)(F)

T = temperature, F

R 1 2 = heat-transfer resistance of the pipe wall, (sec)(F)/Btu

r 1 = pipe inside radius, ft

r 2 = pipe outside radius, ft

k 1 2 = pipe thermal conductivity, Btu/(sec)(ft)(F)

L = pipeline length between points a and b, ft

R 2 3 = heat-transfer resistance of the insulation, (sec)(F)/Btu

r 3 = insulation outside radius, ft

k 2 3 = thermal conductivity of the insulation, Btu/(sec)(ft)(F)


A-2

R 3 4 = heat-transfer resistance of the ground, (sec)(F)/Btu

z = distance of pipe axis below ground surface, ft

k 3 4 = thermal conductivity of the ground, Btu/(sec)(ft)(F)

q = heat transferred through the pipe wall, Btu/sec

T 4 = ground surface temperature, F

To = isothermal pipeline temperature, F

∆T = log mean temperature difference, F

∆ T max = maximum temperature difference in pipeline length L, F

∆ Tm i n = minimum temperature difference in pipeline length L, F

W = LNG weight flow rate, lb/sec

w = specific weight of LNG, lb/ft 3

f = friction factor

P = pump power, hp

e = pump efficiency.

Derivation

For the steady flow of a fluid through any apparatus, the general energy equation
may be written across any two points of interest as follows (Figure A-1):

FIGURE A-1. NOTATION FOR STEADY-FLOW ENERGY EQUATION


A-3

(A-1)

Steady flow, as defined here, is a flow in which the flow characteristics, such as pres-
sure, velocity, density, etc., at a point do not change with time, although they can
change with space variation. Applying the general energy equation to a level LNG pipe-
line, which is subjected to no mechanical work, and which contains a fluid of constant
density (va = vb = vb and Va = Vb) we obtain

(A-2)

or
(A-3)

Equation (A-3) states that the change in LNG internal energy is equal to the sum of the
heat transferred to the LNG from. the surroundings and the energy expended in fluid
friction (pressure drop) over the pipeline length of interest.

The internal energy difference is simply

(A-4)

The heat transferred into the pipeline is a function of pipeline length since the
LNG temperature varies with pipeline length. In general, the heat transferred across
a composite structure may be expressed in terms of a temperature difference and a
summation of heat-transfer resistances as

(A-5)

For the case of the LNG pipeline, these resistances include the fluid film on the inside
wall of the pipe, the pipe wall itself, the pipe insulation, the earth in which the pipe
and insulation are buried, and the air film on the surface of the ground. Two of these,
the fluid film on the inside wall of the pipe and the air film on the surface of the ground,
will be eliminated as follows:

(1) The temperature of the inside pipe-wall surface will be assumed


equal to the LNG temperature

(2) The ground-surface temperature will be specified instead of the


ambient air temperature.

These are both conservative assumptions leading to greater heat influx than will actually
occur. Consequently, the heat-transfer resistances of interest are those of the pipe
wall, the insulation, and the soil surrounding the pipe. These are shown in Figure A-2.
A-5

The heat-transfer resistance of a composite cylinder consisting of pipe wall and


insulation are given by

(A-6)

(A-7)

M c A d a m s ( 4 9 ) gives a relation for the heat transferred from a buried isothermal heat
source to the ground surface. The resistance factor between a horizontal cylindrical
heat sink and a semi-infinite body in which it is encased (ground) is given as

(A-8)

The heat transfer from the ground surface to the buried LNG pipeline, for the case of
an isothermal pipeline, is then

(A-9)

As previously mentioned, the LNG pipeline is not an isothermal heat sink. A


good approximation of the heat transfer in the LNG pipeline can be made by replacing
the constant temperature difference in Equation (A-9) by a suitable mean temperature
difference. The mean temperature difference commonly employed in heat-transfer
situations such as this, known as the log mean temperature difference, is defined in
general terms as

(A-10)

For the pipeline under consideration

(A-11)
A-6

Substituting Equation (A-11) into Equation (A-9) for the temperature difference, the
heat transferred into the pipeline as a function of axial distance along the pipeline is

(A-12)

Finally, the heat transfer per unit weight is given by

(A-13)

The last term of the energy equation, Equation (A-3), is the pipeline pressure
drop. Any fluid-mechanics text gives the pressure drop in terms of a friction factor
and other parameters as

(A-14)

Substituting from Equations (A-4), (A-6), (A-7), (A-8), (A-13), and (A-14) into
Equation (A-3) yields

(A-15)

Solving for the pipeline length L, Equation (A-15) becomes

(A- 16)
A-7 and A-8

Equation (A-16) relates the LNG temperature to the pipeline length, i.e., for an
initial LNG temperature of T a , the temperature will rise to a value T b in a length of
pipe equal to L. For most practical cases the heat-transfer resistance of the pipe wall
is negligible in relation to the other resistances, and Equation (A-16) may be simplified
to

(A-17)

where

and

The mechanical power required to pump LNG through a pipeline length L is given
by

(A-18)

where the pressure drop is obtained from Equation (A-14). The temperature rise of
liquid in the pump due to pumping is in most cases minor for pumps of reasonable effi-
ciency. For example, when all pump losses are assumed to result in a temperature in-
crease, the temperature increase is less than 10 F for a pump-pressure difference of
1000 psi and a low pump efficiency of 50 percent.
This Page Intentionally Left Blank
This Page Intentionally Left Blank
Copyright, 2003
All Rights Reserved by Pipeline Research Council International, Inc.

PRCI Reports are Published by Technical Toolboxes, Inc.

3801 Kirby Drive, Suite 340


Houston, Texas 77098
Tel: 713-630-0505
Fax: 713-630-0560
Email: info@ttoolboxes.com

You might also like