0% found this document useful (0 votes)
415 views188 pages

Backfill Materials For Underground Power Cables Phase 1

Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
0% found this document useful (0 votes)
415 views188 pages

Backfill Materials For Underground Power Cables Phase 1

Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
You are on page 1/ 188

BACKFILL MATERIALS FOR UNDERGROUND POWER CABLES

PHASE 1

ERR I EL-506
(Research Project 7841-1)

Interim Report
Thermal Resistivity Measurement Methods, Backfill Treatments
Heat and Moisture Flow Analysis

June 1977

Prepared by
D
Department of Civil Engineering
UNIVERSITY OF CALIFORNIA, BERKELEY

PRINCIPAL INVESTIGATORS
James K. Mitchell
T-C. Kao
Omar N. Abdel-Hadi

Prepared for

Electric Power Research Institute


3412 Hillview Avenue
Palo Alto, California 94304

EPRI Project Manager


T. J. Rodenbaugh
tp
IS "
distribution of thboocumcnt
DISCLAIMER

This report was prepared as an account of work sponsored by an


agency of the United States Government. Neither the United States
Government nor any agency thereof, nor any of their employees,
makes any warranty, express or implied, or assumes any legal liability
or responsibility for the accuracy, completeness, or usefulness of any
information, apparatus, product, or process disclosed, or represents
that its use would not infringe privately owned rights. Reference
herein to any specific commercial product, process, or service by
trade name, trademark, manufacturer, or otherwise does not
necessarily constitute or imply its endorsement, recommendation, or
favoring by the United States Government or any agency thereof. The
views and opinions of authors expressed herein do not necessarily
state or reflect those of the United States Government or any agency
thereof.

D IS C L A IM E R

Portions of this document may be illegible in electronic image


products. Images are produced from the best available
original document.
LEGAL NOTICE

This report was prepared by the University of California, Berkeley (UCB), as


an account of work sponsored by the Electric Power Research Institute, Inc.
(EPRI). Neither EPRI, members of EPRI, UCB, nor any person acting on behalf
of either: (a) makes any warranty or representation, express or implied, with
respect to the accuracy, completeness, or usefulness of the information con­
tained in this report, or that the use of any information, apparatus, method,
or process disclosed in this report may not infringe privately owned rights;
or (b) assumes any liabilities with respect to the use of, or for damages
resulting from the use of, any information, apparatus, method, or process
disclosed in this report.

I
Abstract

Because the allowable current loading of buried electrical transmission

cables is frequently limited by the maximum permissible temperature of the

cable or of the surrounding ground, there is need for cable backfill materials

that can maintain a low thermal resistivity (less than 50 °C-cm/watt) even

while subjected to high temperatures for prolonged periods. This report des­

cribes the results of studies aimed at development of improved methods for

placing backfill around underground power cable systems and special treat­

ments to reduce the thermal resistivity and increase the thermal stability of

the backfill materials.

The thermal needle method has been selected as the simplest, fastest, and

most reliable method for measurement of the thermal resistivity of backfill

materials in the laboratory. A special thermal needle design has been de­

veloped for this project.

Unlike most mechanical properties of sands and gravels, the thermal con­

ductivity of backfill materials is not affected by the method of compaction

used, provided comparisons are made at a given value of density. On the other

hand, samples compacted wet and then dried have a lower thermal resistivity

than do samples compacted dry.

Additives which can either (1) prevent water migration in the soil due

to temperature gradients, or (2) substitute for water in bridging the gaps

between particles have potential usefulness for backfill treatment. Three

water absorbing polymers were tested and found to retard, but not prevent

water migration. Once dried, samples treated with these materials had poorer

thermal properties than did the untreated soil.

iii
No additives have been found that can produce a material with a thermal

resistivity significantly less than that of the untreated wet compacted mate­

rial, i.ev 30-60 °C-cm/watt, depending on the soil type, density, etc. On

the other hand, portland cement, C-7 (a cement-based additive), inexpensive

waxes, and asphalt treatment can maintain a low thermal resistivity, even

after complete drying of the material. Hence, they act as suitable water sub­

stitutes, are not susceptible to migration, and appear to be durable, although

testing is not yet complete to confirm all aspects of durability.

A finite element computer program HEAT has been used to study transient

and steady-state heat flows and temperature distributions for typical buried

cable systems. This program allows for detailed modeling of geometrical con­

ditions, native soil and backfill properties, surface cover, and weather con­

ditions. The dominating influence of the thermal resistivity of the trench

backfill on the temperature distributions and allowable thermal loading is

readily apparent from the analyses.

Theories for moisture migration under thermal gradients have been studied,

and the Philip and deVries theory has been selected for analysis of trench and

backfill systems. Thermal and moisture diffusion properties of three sands

characteristic of those to be encountered in actual buried cable systems have

been measured. Analyses of moisture flows to be expected for different cable

operating temperatures are planned.

The next phase of this investigation will include field tests to evaluate

both the performance of the most promising backfill treatments and the accu­

racy of the developed heat and moisture flow predictive methods.

iv
Acknowledgments

A number of individuals and organizations have made important contribu­

tions to the success of this project, and their assistance is acknowledged

with thanks.

Kenneth M. Smith, Graduate Research Assistant during 1975-76 contributed

to the theoretical analyses of heat and moisture flow. Kin W. Lee assisted

with laboratory testing.

Herman Halperin, Thomas J. Rodenbaugh, and Ralph W. Samm of the Electric

Power Research Institute provided valuable background information and guidance

to the research.

K. N. Akay of the Pacific Gas and Electric Company provided the initial

stimulation which led to the establishment of this project and provided help­

ful suggestions and guidelines during the progress of the work. Personnel of

the Pacific Gas and Electric Company Research Laboratory in San Ramon, Cali­

fornia contributed useful information.

J. R. Easterling, Los Angeles Department of Water and Power, has parti­

cipated in project review meetings and has provided large samples of the

thermal backfill used for underground installations in Los Angeles.

Samples of ICC Subcommittee "Round Robin Sand" used for thermal resisti­

vity testing by several organizations were provided by Baltimore Gas and

Electric Company.

Dave Lesnini of Chevron Research Corporation has provided technical in­

formation and wax samples for use as a backfill admixture. Samples of C-7

and Chemicolime were provided by Takanaka Komuten Company, Japan, and the

Master Builders Company, Emeryville, California, provided samples of Embaco.

v
TABLE OF CONTENTS

Page

CHAPTER I. INTRODUCTION 1-1

CHAPTER II. MEASUREMENT OF THERMAL RESISTIVITY 2-1

Introduction 2-1
Thermal Needle Method 2-1
Shannon and Wells 2-2
"Rhometer" 2-3
Guarded Hot Plate Method 2-3
Rapid "k" Method 2-8
Conclusions on Resistivity Measurement Methods 2-8
Thermal Needle Design 2-8
Radial Thermal Resistivity 2-14
Specific Heat 2-14
Vertical Thermal Resistivity 2-18
Factors Influencing the ThermalNeedle Method 2-21
Effect of Finite Sample Diameter 2-21
Current Variation During Test 2-22
Boundary Conditions 2-26
Effect of Needle Composition 2-26
Influence of Test Duration 2-28

CHAPTER III. FACTORS CONTROLLING THERMAL RESISTIVITY 3-1

Introduction 3-1
Soil Composition 3-1
Density 3-3
Water Content 3-3
Particle Size and Shape 3-3
Grain Size Distribution 3-6
Temperature 3-6
Effect of Compaction Method (Influence of Sand Fabric) 3-6

vii
PAGE

CHAPTER IV. BACKFILL TREATMENTS 4-1

Introduction 4-1
Treatment of Monterey No. 0 Sand 4-1
Treatment of Fire Valley Thermal Sand 4-4
Effect of Polymers on Thermal Resistivity 4-6
of Fire Valley Thermal Sand
Waxes and Asphalt as Thermal Stabilizers 4-14
Wax Selection and Properties 4-18
Preparation of Wax-Treated Samples 4-18
Effects of Waxes on Fire Valley Thermal Sand 4-19
Effect of Slack Wax on Monterey No. 0 Sand 4-19
and Round-Robin Sand
Emulsified Wax Additives 4-24
Effect of Temperature and Temperature History 4-24
on Wax-Treated Sands
Evaluation of 3M Stabilizer 4-28
Conclusions 4-28

CHAPTER V. TEMPERATURE DISTRIBUTIONS AROUND BURIED 5-1


CABLES AND HEAT FLOW ANALYSES

Introduction 5-1
Previous Solutions 5-2
Discussion of Solutions 5-13
Program HEAT 5-15
Steady State Heat Flow Analyses 5-17
Effect of Trench Size and Configuration 5-23
Effect of Finite Cable Size 5-23
Parameter Study 5-28
Transient Conditions 5-44
Heat Removal By Buried Thermal Conductors 5-49

CHAPTER VI. MOISTURE MIGRATION UNDER THERMAL GRADIENTS 6-1

Introduction 6-1
Statement of the Problem 6-1
Taylor and Cary Theory 6-2
Philip and DeVries Theory 6-3
Vapor Phase Transport 6-4
Liquid Phase Transport 6-5
Interaction of the Vapor and Liquid Phases 6-7
Comprehensive Theory 6-7
Evaluation of the Soil Parameters 6-9
Hydraulic Conductivity, Kq 6-9
Isothermal Water Diffusivity, Dq 6-20
Thermal Water Diffusivity, DT 6-24
Thermal Diffusivity, X 6-28
Planned Application 6-32

CHAPTER VII. SUMMARY AND CONCLUSIONS 7-1


R—1
REFERENCES

viii
LIST OF FIGURES

FIGURE PAGE

1 Original Time Factor Curves for Temperature Change at 2-4


Center of Cylinder

2 Corrected Time Factor Curves for Use with the Shannon 2-5
and Wells Method

3 Cross Section of Rhometer Apparatus 2-6

4 Arrangement for the Guarded Hot Plate Method 2-7

5 Thermal Needle Design No. 1 2-10

6 Comparison of Thermal Resistivity Values Measured in 2-11


Different Laboratories

7 Thermal Needle Design No. 2 2-12

8 UC Thermal Needle 2-13

9 Thermal Needle and Sample Arrangement 2-15

10 Centerline Temperature as a Function of Time for a Sample 2-16


of Dry Monterey No. 0 Sand Initially at 20°C in a Water Bath
at 40°C

11 Time Factor Curve for Radial Heat Flow 2-17

12 Temperature as a Function of Time at Two Points along 2-19


Sample Axis with Conductive Bottom End Plate

13 Time Factor at 50% Temperature Change (Thermocouple No. 1) vs 2-20


Ratio of Thermal Resistivities in Radial and Axial Directions

14 Theoretical Temperature Distributions - Thermal Needle Method 2-23

15 Typical Thermal Needle Test Result for Storage Battery 2-24


Power Supply

16 Thermal Needle Test Result for Constant Current 2-25


Power Supply

17 Influence of Current Variation on Value of Thermal Resistivity 2-27


Determined in Thermal Needle Test

18 Effect of Needle Composition on Results of Thermal Needle Test 2-29

19 Analysis of the Influence of Test Duration on Thermal 2-30


Needle Test Results

ix
FIGURE PAGE

20 The Effect of Density on the Thermal Resistivity of Dry 3-4


Fire Valley Thermal Sand

21 The Influence of Water Content on the Thermal Resistivity 3-5


of Fire Valley Thermal Sand

22 Particle Size Distribution Curves of Backfill Materials 3-7

23 The Temperature Dependence of the Thermal Resistivity of 3-8


Dry Quartz Sand and Water

24 Influence of Method of Sample Preparation on Thermal 3-10


Resistivity (Dry Monterey No. 0 Sand)

25 Influence of Method of Sample Preparation on Thermal 3-11


Resistivity (Saturated Monterey No. 0 Sand)

26 Thermal Resistivity vs. Density for Dry Fire Valley Thermal 3-12
Sand Samples Prepared by Different Methods

27 Thermal Resistivity vs. Density for Moist Fire Valley 3-13


Thermal Sand Samples Prepared by Different Methods

28 Thermal Resistivity vs. Density for Dry Round-Robin Sand 3-14


Samples Prepared by Different Methods

29 Thermal Resistivity vs. Density for Moist Round-Robin Sand 3-15


Samples Prepared by Different Methods

30 Effect of Water During Compaction on the Thermal Resistivity 3-17


of Dry Fire Valley Thermal Sand

31 Effect of C-7 on the Thermal Properties of Monterey No. 0 4-5


Sand During Curing and Drying

32 Effect of Dow Chemical Polymer on Thermal Resistivity and 4-11


Water Content of Fire Valley Sand

33 Effect of General Mills Polymer on Thermal Resistivity 4-12


and Water Content of Fire Valley Sand

34 Effect of Union Carbide Viterra on Thermal Resistivity 4-13


and Water Content of Fire Valley Sand

35 Thermal Resistivity as a Function of Water Content for 4-15


Polymer-Treated Fire Valley Thermal Sand

36 Effect of Wax (CD 150/160) on Thermal Conductivity of 4-17


Fire Valley Thermal Sand

X
FIGURE PAGE

37 Effect of Slack Wax on Thermal Resistivity of Fire Valley 4-20


Thermal Sand

38 Effect of Waxes on Thermal Resistivity of Fire Valley 4-21


Thermal Sand

39 Effect of Slack Wax on Thermal Resistivity of Monterey 4-22


No. 0 Sand

40 Effect of Slack Wax on Thermal Resistivity of Round- 4-23


Robin Sand

41 Effect of Emulsified Wax on Thermal Resistivity of Fire 4-25


Valley Thermal Sand

42 Effect of Temperature on Thermal Resistivity of Slack Wax- 4-26


Treated Fire Valley Thermal Sand

43 Effect of Temperature on Thermal Resistivity of Slack Wax- 4-27


Treated Round-Robin Sand

44 Finite Element Mesh Used for Analysis of Temperature 5-19


Distributions around Buried Power Cable

45 Maximum Heat Input Rate to Limit Cable Sheath Temperature 5-21


to 90°C in a Homogeneous Soil System

46 Allowable Heat Input Rate to Limit the Cable Sheath 5-22


Temperature to 90°C for Different Trench Backfill Resistivities

47 Effect of Natural Ground Instead of Special Backfill in Upper 5-25


Portion of Trench on Allowable Heat Input

48 Influence of Natural Ground Resistivity on the Allowable Heat 5-26


Input to a Cable Trench System

49 Comparison Between Line Heat Source and Cable of Finite Cross 5-27
Section on Allowable Heat Input Rate to Limit Cable Sheath
Temperature to 90°C

50 Effect of Resistivity of Backfill Soil 5-30

51 Effect of Resistivity of Native Soil 5-32

52 Effect of Cable Size 5-33

53 Effect of Trench Depth onAllowable Heat Input 5-34

54 Effect of Cable Depth onAllowable Heat Input 5-35

xi
FIGURE PAGE

55 Effect of Cover Thickness on Allowable Heat Input 5-36

56 Effect of Native Soil Temperature on Allowable Heat Input 5-37

57 Effect of Air Temperature on Allowable Heat Input 5-39

58 Effect of Wind Velocity on Allowable Heat Input 5-40

59 Effect of Solar Radiation on Allowable Heat Input 5-41

60 Effect of Maximum Allowable Cable Surface Temperature 5-42


on Allowable Heat Input

61 Temperature Contours for a Cable at 90°C 5-43

62 Allowable Heat Input Rates for Multi-Cable Systems 5-45

63 Allowable Heat Input Rate for Different Backfill Materials 5-46


in the Top 6" of the Trench

64 Time Required to Reach Indicated Temperature as a Function 5-47


of Heat Input Rate

65 Cable Heating During a Hypothetical Heat Wave 5-48

66 Volumetric Water Content-Suction Head Relationship for 6-14


Fire Valley Thermal Sand

67 Volumetric Water Content-Suction Head Relationship for 6-15


Round-Robin Sand

68 Volumetric Water Content-Suction Head Relationship for 6-16


Monterey No. 0 Sand

69 Variation of Hydraulic Conductivity of Fire Valley 6-17


Thermal Sand with Degree of Saturation

70 Variation of Hydraulic Conductivity of Round-Robin 6-18


Sand with Degree of Saturation

71 Variation of Hydraulic Conductivity of Monterey No. 0 6-19


Sand with Degree of Saturation

72 Variation of Isothermal Liquid Diffusivity of Fire Valley 6-21


Thermal Sand with Degree of Saturation

73 Variation of Isothermal Liquid Diffusivity of Round-Robin 6-22


Sand with Degree of Saturation

74 Variation of Isothermal Liquid Diffusivity of Monterey No. 0 6-23


Sand with Degree of Saturation

xii
FIGURE PAGE

75 Variation of Isothermal Vapor Diffusivity of Fire Valley 6-25


Thermal Sand with Degree of Saturation

76 Variation of Isothermal Vapor Diffusivity of Round-Robin 6-26


Sand with Degree of Saturation

77 Variation of Isothermal Vapor Diffusivity of Monterey No. 0 6-27


Sand with Degree of Saturation

78 Variation of Thermal Water Diffusivity of Fire Valley Thermal 6-29


Sand with Degree of Saturation

79 Variation of Thermal Water Diffusivity of Round-Robin Sand 6-30


with Degree of Saturation

80 Variation of Thermal Water Diffusivity of Monterey No. 0 6-31


Sand with Degree of Saturation

81 Variation of Thermal Diffusivity of Fire Valley Thermal 6-33


Sand with Degree of Saturation

82 Variation of Thermal Diffusivity of Round-Robin Sand 6-34


with Degree of Saturation

83 Variation of Thermal Diffusivity of Monterey No. 0 Sand 6-35


with Degree of Saturation

xi±i
LIST OF TABLES

PAGE

TABLE 1 AVERAGE RESISTIVITY VALUES FOR SOME SOIL 3-2


CONSTITUENTS AND ALLIED MATERIALS

TABLE 2 EFFECT OF ADDITIVES ON THERMAL RESISTIVITY OF 4-2


MONTEREY NO. 0 SAND

TABLE 3 EFFECT OF ADDITIVES ON THERMAL RESISTIVITY OF 4-7


FIRE VALLEY THERMAL SAND

TABLE 4 EFFECT OF WAX ON THERMAL RESISTIVITY OF 4-16


FIRE VALLEY THERMAL SAND

TABLE 5 THERMAL CONDUCTIVITY VALUES OF SEVERAL BACKFILL- 4-29


ADDITIVE SYSTEMS

TABLE 6 ALLOWABLE HEAT INPUT RATE FOR DIFFERENT 5-24


TRENCH CONFIGURATIONS

TABLE 7 HEAT REMOVAL BY BURIED THERMAL CONDUCTORS 5-51

xiv
I. INTRODUCTION

The allowable current loading of buried electrical transmission cables is

frequently limited by the maximum permissable temperature of the cable or the

surrounding ground. Temperatures greater than 50° to 60° C. may lead to

breakdown of cable insulation and thermal runaway if the surrounding cable

trench backfill is unable to conduct the heat away as rapidly as it is gen­

erated. The problem is aggravated by the fact that the moisture in most

backfill materials used around buried cables migrates away from the cable

resulting in marked increase in thermal resistivity. New cable systems are

under development that are designed for most efficient operation at tempera­

tures as high as 90° C.

Hence, there is a real need for cable backfill materials that can main­

tain a low thermal resistivity (less than 50° C-cm/watt if possible), even

while subjected to high temperatures for prolonged periods. The maximum

operating temperature would then become independent of the surrounding back­

fill and soil conditions and would be controlled by such factors as insulation

and internal stresses. The accompanying benefits would include a substantial

increase in allowable cable loading and reduced maintenance, with an attendant

decrease in the cost of underground transmission installations.

The objective of this project is to develop improved methods for placing

backfill around underground power cable systems and special treatments to

reduce the thermal resistivity and increase the thermal stability of the

backfill materials. The scope of work during the two years since initiation

of research on this project has included several phases and tasks aimed at

meeting this objective. Following an initial period used to evaluate the

present states of knowledge and practice, work has proceeded concurrently

1-1
along both experimental and theoretical lines. Specific topics have in­

cluded the following:

1. Evaluation of methods for measurements of the thermal resistivity

of backfill materials in the laboratory.

2. Development of improved and standardized procedures for laboratory

measurement of thermal resistivity.

3. Evaluation of factors controlling the thermal resistivity of soils.

4. Evaluation of the influences of backfill compaction method on thermal

resistivity.

5. Experimental studies of the effectiveness of a large number of

additives and backfill treatment methods for reducing and main­

taining a low thermal resistivity.

6. Theoretical analyses of temperature distributions and heat flows

around buried power cables.

7. Theoretical study of the moisture flows in partly saturated backfill

systems under thermal gradients.

8. Measurement of the parameters necessary for prediction of moisture

flow under thermal gradients for typical backfill materials.

The accomplishments and conclusions resulting from each of these studies

are presented in this report.


II. MEASUREMENT OP THERMAL RESISTIVITY

Introduction

Several methods have been used for evaluation in the laboratory of the

thermal resistivity of soils of the type used as backfill materials around

buried power cables. A thorough study of each of these has been made in

order to establish their advantages and disadvantages and to enable selection

of the most suitable procedures for use in this project.

Two methods; namely, the thermal needle method and the Shannon and

Wells method have been used most extensively. In addition a "Rohmeter"

method, a guarded hot plate method and a "Rapid k" method have been developed.

Each of these methods is described below.

Thermal Needle Method

The thermal needle method is based on the measurement of the rate of

temperature rise along a line heat source within an infinite, homogenous

medium. The temperature 0 at any time t at the heat source will be

(Carlslaw and Jaeger, 1959).

(1)

where Q = heat input per unit length per unit time

k = thermal conductivity = 1/p

p = thermal resistivity

4at

a = thermal diffusivity

r = radial distance from heat source

2-1
For small values of X and larger values of t equation (1) becomes:

( 2)

Therefore

(3)
A0 = iik (A ln fc) = 4fF p (A ln t)

Thus if heat is applied at a known constant rate to a line embedded in the

medium of interest and the temperature of the line is measured as a function

of time, a straight line relationship in indicated between 0 and In t, with

a slope that is proportional to the thermal resistivity, p. In practice the

line heat source is approximated by a small diameter needle.

The thermal needle method involves relatively simple apparatus and in­

strumentation, and probes can be fabricated for both laboratory and field

use. It has a great advantage in that the thermal resistivity can be com­

puted directly from the test data without knowledge of the heat capacity of

the material. This method has been used successfully by a number of elec­

trical utilities. Further details concerning this method are presented later

in this chapter.

Shannon and Wells (Shannon and Wells, 1959)

In this method the thermal diffusivity of a cylindrical sample is eval­

uated by measuring the temperature change at the center of a warm sample

(40° C.i) after immersion into a colder water bath (20° C.*). The thermal

resistivity is computed from the measured diffusivity based on the assumption

of a constant specific heat, C. A value of C equal to 0.2 cal/gm/°C for the

soil solids has been suggested. This method suffers from two disadvantages.

1. The assumption of specific heat equal to 0.2 cal/gm/°C can induce an

error as high as 15% for the sands we have tested in this project. «
2-2
Separate measurement of specific heat is required for best results.

2. The Time Factor Curve for temperature change at the center of a

cylinder of diameter D and height 2D presented by Shannon and Wells

was found to be in error. Fig. 1 shows the original theoretical

solution given by Shannon and Wells. Reanalysis (Fig. 2), shows

curve 0 in Fig. 1 to be in error by about 20 percent. Values of

thermal resistivity determined using this curve will therefore be

too high by 20 percent.

"Rhometer" (J. Stolpe, 1969)

The rhometer method involves application of steady-state radial heat

flow theory. Measurement of the thermal resistivity of a soil by this method

requires determination of the temperature difference between two concentric

cylinders, one of which serves as a heat source and the other serves as a

heat sink, as shown in Fig. 3. In order to achieve the steady state, "the

rhometer must be allowed to run for about ten hours before making measure­

ment" (Stolpe, 1969). The long time required for the test is a disadvantage

in itself. In addition, some moisture migration is probable during a test

of this duration, further complicating the measurement.

Guarded Hot Plate Method (A.S.T.M. Designation: C177-71)

This method also requires steady state heat flow. The test arrangement

is shown as Fig. 4. Two samples are required for each test. Thermal

resistivities are interpreted from the temperature difference between the

heater and the cold plates. This method is time-consuming, and water

migration during the test may occur.

2-3
t | I i I i r
Notation
/u -Percent Temperature Change
6f)'0r
-
C Volumetric Heat Capacity,
-
Btu./cuft/°F
X - Unit Dry Weight, tp/cu ft.
C, - Specific Heat of Dry Soil
W = Water Content, % Dry Weight
k : Thermal Conductivity, Btu/ft/ht/P
% Temperature Change,

t1 Time .Hours
OfInitial External and Interna! Temp
at O
e, -Uniform External Temperature
Suddenly Applied
9C -Temp, at Center of Cylinder at Time t
otp-O, oc-%-, atyt-iOO%,ec-6,-
D -Diameter of Cylinder, feet
H -1/2 Cylinder Length
T - Dimensionless Time Factor

_ ^(2) H-CD
_ Equations © HS2D
100-
(a) Thermal Diffusivity, a s ^T/r
(b) Volumetric Heat Capacity (Unfrozen Soil), C -X(C, + W/foo)
(c) Volumetric Heat Capacity (Frozen Soil), C=x(C/t- 0.5W/)00)
(d) Thermal Conductivity, k-aC
i i I i i I__ I
OD! 0.1
Time Factor, T

FIG. 1 ORIGINAL TIME FACTOR CURVES FOR TEMPERATURE


CHANGE AT CENTER OF CYLINDER
(Shannon and Wells, 1947)

2-4
80 -

H = a>

Time Factor, T

FIG. 2 CORRECTED TIME FACTOR CURVES FOR USE


WITH THE SHANNON AND WELLS METHOD

2-5
Concrete Cylinder
Copper Constonton Thermocouples
Extra-heavy Wall Copper Pipe
Power Resistor
Foam Disk

Plywood Bose
Reinforced Foam Disk
Epoxy Cement

FIG. 3 CROSS SECTION OF RHOMETER APPARATUS

(J. Stolpe, 1969)


Hooks

FIG. 4 ARRANGEMENT FOR THE GUARDED HOT PLATE METHOD

2-7
Rapid "k" Method

This method uses a special apparatus fabricated by "Dynatech". The

sample is sandwiched between a hot and a cold plate. Equilibrium temperatures

at these two plates and the heat flow in the system are used to interpret the

thermal resistivity of the sample. There are several disadvantages in using

this apparatus for the materials and test conditions of interest in this

study:

1. The apparatus is expensive.

2. The sample height must be less them 50 mm, which could give mis­

leading results for coarse grained backfill materials.

3. The operating range of this apparatus for resistivities is in the

range of 200 to 5,000 °C-cm/watt. However, very often the thermal

resistivity of backfill materials is lower than 200 °C-cm/watt.

Conclusions on Resistivity Measurement Methods

Of the five methods for measuring thermal resistivity, the transient

methods (Shannon and Wells method and thermal needle method) are most suitable

for backfill materials because of their relative simplicity and the short time

required for a measurement. By carefully controlling the test details, these

methods give very consistent and reproducible results. Of the two, the

thermal needle method has been found the simplest and offers the added advan­

tage that knowledge of the specific heat is not required for calculation of

resistivity. Consequently, the thermal needle method has been evaluated in

some detail.

Thermal Needle Design

The thermal needles used for measurement of thermal resistivity on this

project were fabricated in the laboratory. The first design is shown in

2-8
Fig. 5. Results obtained with this needle were generally satisfactory as

shown by the comparisons in Fig. 6; however, some dependence of measured

resistivity on current input to the needle was observed. Excessive heat

transfer from the back in the thermocouple plug was at least partly respon­

sible for this.

A second needle design was developed, Fig. 7, in which the length of

resisting wire in the thermocouple plug was minimized. In addition manganin

wire was used for the heating element in place of monel. Manganin has a

higher thermal resistivity than monel, and its thermal resistivity is less

temperature-dependent; thus smaller currents can be used. In addition

manganin is factory insulated. In this needle design the thermocouple wires

were used as spacers between the manganin heating wires.

A third design was made which offers several advantages over the first

two. This needle, referred to as the UC Thermal Needle, is shown schemati­

cally in Fig. 8. The essential features are:

1. A ceramic tube is used instead of stainless steel. Thus the resis­

tivity of the needle itself more closely matches the thermal resis­

tivity of the samples tested.

2. Two thermocouples are used, thus enabling temperature measurement

at two points along the length of the sample.

This needle can be used both as an active heat source, for use in the

thermal needle method, and as a passive heat sensor, for use in the Shannon

and Wells method.

The design of the UC needle makes it possible to measure on the same

sample:

1. thermal resistivity in a vertical direction

2. thermal resistivity in a radial direction

2-9
T-type thermocouple plug

■#26 Monel wire

O #30 T-type thermocouple

Thermocouple
Junctions

Hypodermic needle

63.5 mm 00 = 1.83 mm (0.072")


10=1.37 mm (0.054")

121.9 mm

Epoxy filled

FIG. 5 THERMAL NEEDLE DESIGN NO. 1

2-10
90<ir Insulated Conductors Committee
Project 12-34
Soil Thermal Resistivity Results
Thermal Resistivity-°C-cm/watt from Round- Robin Tests
Thermal Resistivity of Quartz Sand
Nominal Dry Density - 21.2 kN/m3 (135 pcf)

• Thermal needle No.I


o Thermal needle No. 2
A Shannon and Wells Method

-Los Angeles
Water and Power Southern California
Edison

A Ontario Hydro
[^-Georgia Power Co.
Long Island-
Lighting Co.

Percent Moisture (% of dry weight)

FIG. 6 COMPARISON OF THERMAL RESISTIVITY VALUES


MEASURED IN DIFFERENT LABORATORIES
T-type thermocouple plug

n n /

FIG. 7 THERMAL NEEDLE DESIGN NO. 2


Copper wire

Thermocouple /
(T type)

Thermocouple 2
( T type )

Heat Source Wire


(Mangan in *30)

x Thermocouple junction

•-------- Epoxy filled

*------- Ceramic tubing


3l75mm(.i25 H)0.D, 1.575mm (062") I.D.
H3.56 mm
(4.471") Thermocouple 2

56.77mm Thermocouple /

FIG. 8 UC THERMAL NEEDLE


3. the specific heat of the sample.

The methods and nature of the results are best illustrated by an example.

Radial Thermal Resistivity. A sample of dry Monterey No. 0 sand was compacted
O
by pluviation* to a density of 1.696 gm/cm in a standard mold with a thermal

needle at its center as shown in Fig. 9. Polyurethane foam end plates were

glued to the top and botton as shown to prevent heat flow in the vertical

direction. After equilibration of the sample at 20° C in a water bath, the

thermal resistivity in a radial direction was measured by the usual thermal

needle method. A value of 324°C-cm/watt was obtained.

Specific Heat. Temperature equilibrium at 20° C was reestablished, and the

sample was placed in a 40° C water bath. Temperature at the center was

measured as a function of time with the result shown in Fig. 10. The theo­

retical relationship between percent temperature change at the center and

dimensionless time factor T for radial heat flow for these boundary condi­

tions is shown in Fig. 11. The specific heat C could be computed with the

aid of the following relationships:


2
D T,
50
a (4)
r t
50

(5)

where a^ = thermal diffusivity

t = time at 50% temperature change at center of sample

Tj.q = time factor for 50% temperature change (from Fig. 11)

*Pouring the dry sand through air under controlled conditions.

2-14
Heating wire

Thermocouple
Polyurethane foam
end plate.

— Standard
compaction mold

UC Thermal needle

Polyethylene foam
end plate

FIG. 9 THERMAL NEEDLE AND SAMPLE ARRANGEMENT

2-15
1000

5 uoo

$ 1200

O Thermocouple /
A Thermocouple 2

Time (min)

FIG. 10 CENTERLINE TEMPERATURE AS A FUNCTION OF TIME FOR A SAMPLE OF DRY


MONTEREY NO. 0 SAND INITIALLY AT 20°C PLACED IN A WATER BATH AT 40°C
Pgr Cant Temperature Change

FIG. 11 TIME FACTOR CURVE FOR RADIAL HEAT FLOW


= radial thermal resistivity

w = water content of sample

Yd = dry unit weight of sample

D = sample diameter.

For the data shown in Fig. 10 a value of specific heat of 0.18 cal/°C-gm

for the dry sand was calculated. This agrees well with the known specific

heats of quartz and feldspar, which are the major constituents of Monterey

sand, of 0.17 to 0.19 cal/°C-gm.

Vertical Thermal Resistivity. To measure the thermal resistivity in a verti­

cal (axial) direction the bottom insulating end plate was replaced by a 0.25

inch thick stainless steel plate. The sample was then equilibrated at 20° C

and placed in a 40° C water bath and the temperature measured at each thermo­

couple as a function of time, giving the results shown in Fig. 12.

The theoretical relationship in Fig. 13 can be used to determine the

ratio of thermal resistivity in a radial direction to that in the vertical

direction p /p . T is determined from


r z c
a t
T = (6)
C D2

where t is the time for 50% temperature change for thermocouple No. 1.

For the present example the data yield Pr/Pz = 1.02. This indicates

that the thermal properties of Monterey No. 0 sand placed by pluviation are

nearly isotropic. This is significant, since other tests have shown that

grain orientations in samples prepared in this way are decidedly anisotropic,

with preferred orientation of long axes in the horizontal direction.

To illustrate the applicability of the UC thermal needle in the Shannon

and Wells method a test was done using the needle as the passive temperature

2-18
900

K)00—

^ uoo
Tem perature

1200 -

O Thermocouple /
1300-
A Thermocouple 2

UC Thermo! Needle

1400
20 40
Time (min)

FIG. 12 TEMPERATURE AS A FUNCTION OF TIME AT TWO POINTS


ALONG SAMPLE AXIS WITH CONDUCTIVE BOTTOM END PLATE

2-19
_ 50%
temperature
change

FIG. 13 TIME FACTOR AT 50% TEMPERATURE CHANGE


(THERMOCOUPLE NO. 1) VS RATIO OF THERMAL
RESISTIVITIES IN RADIAL AND AXIAL
DIRECTIONS

2-20
sensor for determination of the thermal resistance in radial and axial direc­

tions in a dry sample of Round Robin Sand at a density of 2.14 ^"^cc prepared

by tamping compaction.

In the first part of the test the ends of the sample were insulated

using a 50.8 mm thick layer of styrofoam to insure radial heat flow. After

equilibration of the sample at 40° C, it was placed in a water bath at 20° C,

and the temperature change as a function of time was measured. From the data

a value of Pr equal to 96° C-cm/watt for heat flow in the radial direction was

obtained.

In the second part of the test steel end plates were used instead of

styrofoam and the rate of temperature drop from 40° C to 20° C at the thermo­

couples was again determined. From the theoretical solutions for rate of

temperature change as a function of Pr/Pz: e.g.. Fig. 12, it was determined

that the thermal resistivity in the axial direction p was 113° C-cm/watt;
z
i.e., the sample, prepared by tamping compaction was anisotropic with respect

to thermal resistivity.

Factors Influencing the Thermal Needle Method

An investigation was made into the factors influencing the accuracy of

the thermal needle method. The theoretical analyses were done using finite

element heat flow computer program HEAT (Taylor, 1975). Most of these anal­

yses were done relative to Thermal Needle Design No. 2; however, the conclu­

sions are equally applicable to results obtained using the UC Thermal Needle.

Effect of Finite Sample Diameter. To test the suitability of the analytical

solution (Carlslaw and Jaeger, 1959) for temperature as a function of time

for a constant line heat source in an infinite conducting medium as a basis

for determination of thermal resistivitv bv the thermal needle method, the

2-21
following comparison was made. The theoretical analytical solution assumes

a cylinder of infinite diameter; whereas, in the thermal needle test a sam­

ple of finite diameter (101.6 mm in our studies) is used. Fig. 14 shows the

computed temperature increase as a function of radial distance from the center

for a material having thermal resistivity p equal to 92° C-cm/watt, density


a
of 2.14 gm/cm , and water content of 0 percent, at a time of 500 sec. after

the start of the test.

It may be seen that the theoretical solution for an infinite conducting

medium, the finite element solution for a 762 mm diameter specimen and for a

101.6 mm diameter sample all give the same result. This indicates that

(1) the finite element solution approximates the analytical solution closely,

and (2) the analytical solution, based on assumption of an infinite medium,

commonly used for interpretation of thermal needle tests is valid, at least

up to reasonable testing times, for samples of limited diameter.

Current Variation During Test. A 6 or 12 volt storage battery is often used

as a power supply for the thermal needle test. Small (1-2%) variations in

current output may occur and may impair seriously the accuracy of the test

results.

Fig. 15 shows temperature of thermal needle as a function of time using

a battery power source for a test on the Round Robin Sand used in Insulated

Conductors Committee Project 12-34. The current at each time is indicated

for each data point. The value of p deduced from the data so obtained was

77 thermal ohms. Fig. 16 shows the result for the same specimen using a

constant current input of 2.25 amp. In this case a value of p equal to 91,

some 18 percent greater, was obtained. When the constant current power

supply was programmed to give the same current-time relationship as shown

2-22
50 T T

Sample Oia,
FEM ‘HEAT* Output mm
a 101.6 (constant T outside)
O 762 (insulated outside)

Theoretical solution (a> diameter)


t = 500 sec
p =92 “C-cm/wott
y - 2.14 gm/cc
-23

Water content =0%

2.0 3.0 4.0


Radial Distance from Center (cm)
Fin. 14 THEORETICAL TEMPERATURE DISTRIBUTIONS - THERMAL NEEDLE METHOD
140 l-TJ

Round Robin Sand


Water content = 0% 2.23
Dry density = 2.14 gm/cc 2.23,

IQVe = 2.25 Amp 2.25,


— 50

T e m p e ra tu re (°C )
p - 77 ^“cm/watt 2.22
120
2.25_
1=2.28,
to
I
ro
4^ 0)
Q.
E
£

1=2.3,
100

I i i I I 1 I I I I I l
100 1000
Time (sec)

FIG. 15 TYPICAL THERMAL NEEDLE TEST RESULT FOR STORAGE BATTERY POWER SUPPLY
°
()
C
Temperoture
M
I
NJ
(Ji

t (sec)

FIG. 16 THERMAL NEEDLE TEST RESULT FOR CONSTANT CURRENT POWER SUPPLY
in Fig. 15, a value of 79 ohms was obtained.

A number of additional tests were done to further investigate these

effects. The results are summarized in Fig. 17. An approximate theoreti­

cal analysis of these results has been made which can account for them

reasonably in quantitative terms. The important practical conclusion, how­

ever, is that even small (1 percent) variations in current supply during the

thermal needle test can result in significant errors in the value of computed

thermal resistivity.

Boundary Conditions. The thermal needle test is usually done in the labora­

tory on a confined sample of limited dimensions (101.6 mm diameter by 115.3 mm

high cylinder for our studies). It was desired to determine whether the con­

ditions at the surface of the sample had an influence on the results.

Two tests were done using a sample having a thermal resistivity of 115

thermal ohms. In each case the sample was contained in a stainless steel

cylindrical mold of 6.4 mm thickness. In the first test the sample was

placed in a constant temperature water bath at 20° C, (highly conductive

medium). In the second, the sample was placed in air at 20° C (insulating

medium). Essentially identical needle temperature vs. time curves were

obtained in each case for measurement periods up to 600 seconds.

Analysis by Program HEAT showed that even for values of p as low as 23

the needle temperature will be uninfluenced by the external boundary condi­

tions for times up to 400 seconds, again assuming a 101.6 mm sample diameter.

Effect of Needle Composition. The theory used for the determination of thermal

resistivity from the results of a thermal needle experiment assumes that the

needle is of the same resistivity as the material being tested. As most

needles are constructed using a hollow stainless steel sheath, which has a

2-26
Round Robin Sand

Water content = 0%
60- Dry density s 2.14 gm/cc

Initiol current = 1.90 amp

- 1-2 +1 0 -I -2
Current Variation During Test ( % of Initial Value)

FIG. 17 INFLUENCE OF CURRENT VARIATION ON VALUE OF THERMAL


RESISTIVITY DETERMINED IN THERMAL NEEDLE TEST

2-27
very low resistivity, surrounding the heat source, an analysis using Program

HEAT was made to determine if the results were influenced.

Fig. 18 shows that the presence of the stainless steel does not influ­

ence the rate of temperature increase at the center of a 101.6 mm diameter

sample having a p of 93 thermal ohms. Consequently, the value of p computed

from the results of such a test would not be influenced bv the steel layer.

It may be seen, however, that the actual temperature at the center is less

with steel layer in the system. This is because of the higher heat conducti­

vity away from the center.

Influence of Test Duration. From the results of many measurements by the

thermal needle method it is known that a change in the slope of the tempera­

ture versus log time plot may occur at some stage in the test. This change

was suspected to result from the influence of a finite sample size. Once

the heating front reaches the sample boundaries the actual thermal conditions

differ from those assumed for the test.

Fig. 19 shows the theoretical variation of temperature with time, com­

puted using Program HEAT, for a 101.6 mm diameter sample having a thermal

resistivity of 23 °C-cm/watt. A break in the curve may be seen at t = 400

sec., which reflects the influence of the insulating medium (air) beyond the

cylinder wall. For this case any computation of p from temperature values

after 400 sec. would be in error. As the value of p used for this analysis

was very low (23) the result represents a conservative estimate of the maxi­

mum allowable testing time for samples of this size. The break in the curve

would occur at greater values of time for any material having a higher thermal

resistivity or larger diameter.

2-28
Temperoture at Center of Cylinder (°C )

FIG. 18 EFFECT OF NEEDLE COMPOSITION ON RESULTS OF THERMAL NEEDLE TEST


Tem perature a t C enter of Sam ple (°C )

Slope correctly reflects thermal resistivity

Slope incorrectly reflects

Values from FEM “HEAT" thermal resistivity

« 23°C cm/watt

/steel “ 2-2°C cm/wot,


Insulated Cylinder
101.6mm Dio. Mold

100
t (sec)

FIG. 19 ANALYSIS OF THE INFLUENCE OF TEST DURATION ON THERMAL NEEDLE TEST RESULTS
III. FACTORS CONTROLLING THERMAL RESISTIVITY

Introduction

The major objectives of this project are to develop additives or treat­

ment methods that can be used to reduce the thermal resistivity of backfill

material and to maintain a low thermal resistivity over the service life of

a buried cable system. Understanding what factors determine the thermal

resistivity of a soil is useful to the pursuit of this objective. They

include:

(1) Soil composition.

(2) Density.

(3) Water content.

(4) Particle shape and particle size.

(5) Particle size distribution.

(6) Temperature.

(7) Method of compaction (fabric).

The influences of each of these factors on the thermal resistivity is

summarized in this section.

Soil Composition

Soil is a mixture of water, air, mineral soil grains and sometimes,

organic materials. Table 1 lists the thermal resistivity of the most common

soil constituents and related substances under normal conditions. Of the

listed minerals, quartz has the lowest thermal resistivity and mica has the

highest thermal resistivity. It is obvious from Table 1 that to obtain mini­

mum thermal resistivity, the soil mass should contain as much mineral solids

as possible as the resistivities of most minerals are significantly less

than that of water and air.

3-1
TABLE 1

AVERAGE RESISTIVITY VALUES FOR SOME SOIL


CONSTITUENTS AND ALLIED MATERIALS

Thermal Resistivity
Material (°C - cm/watt)

Quartz | | 7.9
Quartz ± 14.9
Quartz, random orientation 11.0
Quartz glass 79.0
Granite 26-58
CaC03 1 26.3
Marble 34-48
Limestone, dense 45
Ice 45
Sandstone 50
Dolomite 58
Slate 67
Water 165
Mica J_ 170
Pine wood || 265
Pine wood j. 608
Organic Material wet 400
Organic Material dry 700
Air 4000

(Winterkorn, 1960)

3-2
Density

Density is one of the most important factors affecting the thermal resis­

tivity of a backfill material. Fig. 20 shows the relationship between the

thermal resistivity and density of dry Fire Valley Thermal Sand.*

Water Content

A small amount of water added to a dry soil will distribute itself as a

thin film around the soil particles. These films can substantially increase

the contact areas between particles; and, therefore, greatly reduce the ther­

mal resistivity of the soil. The amount of water exceeding that required to

form the film of water around particles fills the voids between particles.

The additional reduction in thermal resistivity due to this additional water

is minimal, because heat transfer is mainly through the low thermal resisti­

vity solid particles and their contacts rather than through pore water.

Fig. 21 shows the relatioship between the thermal resistivity and water con­

tent of Fire Valley Thermal Sand.

Particle Size and Shape

In a study of the effects of particle size and particle shape on the

thermal resistivity of soils, Wiseman and Burrell (1960) concluded that:

A. "In one series of laboratory tests particles of 1 mm and larger

gave resistivities around 200, rising with decreasing particle size

to around 300 when finely fragmented."

B. "Block-shaped" particles gave lower soil resistivities than rounded

particles."

Such conclusions may be of limited validity in general, however,

*Fire Valley Thermal Sand is used as a buried cable backfill material by the
Los Angeles Department of Water and Power.

3-3
Dry Density (kN/m^)
IS.5 20.0 20.5 21.0

Dry Density (pcf)

FIG. 20 THE EFFECT OF DENSITY ON THE THERMAL


RESISTIVITY OF DRY FIRE VALLEY THERMAL SAND

3-4
r so

Dry Density - IS. 6 kN/m° (125 pcf)

Water Content (%)

FIG. 21 THE INFLUENCE OF WATER CONTENT ON THE


THERMAL RESISTIVITY OF FIRE VALLEY
THERMAL SAND

3-5
unless comparisons are made for materials of comparable gradation, mineralogi-

cal composition and density. In most instances differing particle sizes and

shapes give different compacted densities.

Grain Size Distribution

In order to achieve high density, a soil must consist of a range of

particle sizes so that small particles can fit into the voids between large

particles. Round-Robin Sand and Fire Valley Thermal Sand are two sands which

have been widely used as cable backfill materials because of their good thermal

conducting properties. Gradation curves for these materials are shown in

Fig. 22. Also shown is a theoretical curve for a gradation yielding a maxi­

mum density. This curve is defined by:

p = (d/D)°‘5 x 100% (7)

where p is the percent by weight finer than size d, and D is the maximum

particle size (10 mm for this case). Both the Round-Robin sand and the Fire

Valley Sand have gradations that give a high density relative to the theoreti­

cal maximum.

Temperature

Temperature influences the thermal resistivity of soils by two effects.

First, high temperature will dry or drive away the moisture in the soil.

Secondly, the thermal resistivity of each individual soil constituent may be

temperature dependent. Fig. 23 shows the temperature dependence of the ther­

mal resistivity of dry quartz sand and water.

Effect of Compaction Method (Influence of Sand Fabric)

Other research in our laboratory and elsewhere has established that the

method of preparation can have major effects on the mechanical properties

3-6
O Round-Robin sand
t/2
□ p = (d/D) D-/Omm
A Fire Volley thermal sand
P e rc e n t F in e r by W eight

j__L-l ..i..!

d (mm)

FIG. 22 PARTICLE SIZE DISTRIBUTION CURVES OF BACKFILL MATERIALS


Dry Quartz Sand
Unit Weight - 22.0 kN/m 3
(140 pcf)

Water

S 140

O 60
Temperature (°C)

FIG. 23 THE TEMPERATURE DEPENDENCE OF THE THERMAL


RESISTIVITY OF DRY QUARTZ SAND AND WATER

3-8
(e.g., strength, compressibility, deformation modulus) of sand samples pre­

pared to the same density (Mitchell, 1976). These phenomena have been shown

to be related to the different fabrics produced by the different methods;

that ds, different compaction methods produce different grain and inter­

particle contact orientations.

It was considered that different compaction methods might also have signi­

ficant influence on thermal resistivity. If so, there might be good reason to

specify certain field procedures to obtain the most favorable thermal conditions.

Hence, four methods of sample preparation were investigated relative to their

influence on the thermal resistivity of Monterey No. 0 sand (a uniform medium

gravity sand).

1. Pluviation (dry sand was rained into the mold)

2. Moist tamping at 8 percent water content

3. Dry rodding

4. Vibration.

The thermal resistivity of dry and saturated samples was determined by

the thermal needle method. Results are shown in Figs. 24 and 25 in the form

of thermal resistivity p versus relative density.*

Samples of Round-Robin Sand and Fire Valley Thermal Sand were also com­

pacted by tamping, vibration in one layer, vibration in five layers, and

rodding. Thermal resistivities were determined by the thermal needle method.

Results are shown in Figs. 26 and 27 for samples of Fire Valley Thermal Sand

prepared dry and moist, respectively. Figs. 28 and 29 give the results for

the Round-Robin Sand.

e - e
♦Relative density, D , is defined as: D = ----------- where e , e . ,
r r e - e . max min
max mm
and e are the maximum, minimum and actual void ratios, respectively.

3-9
Dry Density (k N/m^)
15.5 16.0 !6.5 170
“1 I I T
Dry Density (pcf)

O Ptuviated dry
▲ Moist tamped, then oven-dried (i40eC)
V Dry Rodding
□ Vibrated dry

350 —

0 70 8t
Relative Density (%)

’IG. 24 INFLUENCE OF METHOD OF SAMPLE PREPARATION ON THERMAL RESISTIVITY


(DRY MONTEREY NO. 0 SAND)
Dry Density (kN/m3)

O Ptuviated
A Moist tamped
O Vibrated
(°C-cm/wQtt)
p

40 50 60 70 80 90
Relative Density (%)

FIG. 25 INFLUENCE OF METHOD OF SAMPLE PREPARATION ON


THERMAL RESISTIVITY (SATURATED MONTEREY NO. 0
SAND)

3-11
Dry Density (kN/m3)

a Tamped
O Vibrated (4-25.4 mm layers, I-12mm layer)—
v Rodded

» 135
Dry Density (pcf)

FIG. 26 THERMAL RESISTIVITY VS. DENSITY FOR DRY FIRE VALLEY


THERMAL SAND SAMPLES PREPARED BY DIFFERENT METHODS

3-12
Moist Density (kN/rrt3)

A Tamped
■ Vibrated (/ -//3.6mm /oyer)
O vibrated (4 -25.4 mm layers, 1-12mm layer)
20 — v Rodded —

6% Water Content

/Oi______ I__________ I__________ I__________ I__________ I__________


125 130 135 140 145 150
Moist Density (pcf)

FIG. 27 THERMAL RESISTIVITY VS. DENSITY FOR MOIST FIRE VALLEY


THERMAL SAND SAMPLES PREPARED BY DIFFERENT METHODS
Dry Density (k N/m3)

A Tamped
□ Vitro ted (4-25.4mm layers, 1-12mm layer)
v Rodded
( °C- cm/Wott)
p

135
Dry Density (pcf)

FIG. 28 THERMAL RESISTIVITY VS. DENSITY FOR DRY ROUND-ROBIN


SAND SAMPLES PREPARED BY DIFFERENT METHODS

3-14
Moist Density (kN/m?)

A Tamped
□ Vibroted (4-25.4 mm layers, I~l2mm layer)
V Rodded

6 % Water Content

135
Moist Density (pcf)

FIG. 29 THERMAL RESISTIVITY VS. DENSITY FOR MOIST ROUND-ROBIN


SAND SAMPLES PREPARED BY DIFFERENT METHODS

3-15
These results show that, as was found for Monterey Sand, the method of

sample preparation had no significant effect on thermal resistivity as com­

pacted. It is reasonable to conclude, therefore, that the thermal resistivity

of backfills in the field should be essentially independent of the placement

procedure used.

On the other hand the test results did suggest that the thermal resisti­

vity of a sample placed dry might be significantly greater than for the same

soil compacted wet and then dried. This phenomenon has been studied further

using Fire Valley Thermal Sand, which is a mixture of quartz sand (35 percent)

and 6.5 mm (1/4 inch) limestone aggregate (65 percent). Shown in Fig. 30 are

relationships between thermal resistivity and dry density for samples prepared

dry and for samples compacted by (1) moist tamping at 6 percent water content

and then dried, (2) moist vibration at 6 percent water content and then dried,

and (3) moist vibration at 8 percent water content and then dried. The latter

series of tests was done by the Los Angeles Department of Water and Power. It

may be seen that the thermal resistivity of samples compacted wet and then

dried are substantially less than for samples compacted dry.

It appears that this difference arises from two causes:

1. Because of the bimodal grain size distribution curve for the sand

(see Fig. 22), there is segregation of particle sizes when dry com­

paction is used. This leads to variable fines distribution within

the sample and increased resistivity.

2. Slight solution of the limestone aggregate in the water of compac­

tion provides precipitates at interparticle contacts during drying.

This produces an enlarged solid interparticle contact zone thus de­

creasing thermal resistivity.

3-16
Dry Density (kN/m3)
20.0

Compacted dry
( C-cm/watt)

Compacted wet
p
Thermal Resistivity,

Moist tamped ot 6% water content, then dried(60°C)


Moist vibrated ot 6% water content, then dried (60°C)
Moist vibrated at 8% water content, then dried (60°C)
(Tests by Los Angeles Department of Water and —
Power)

130
Dry Density (pcf)

FIG. 30 EFFECT OF WATER DURING COMPACTION ON THE THERMAL


RESISTIVITY OF DRY FIRE VALLEY THERMAL SAND

3-17
The practical consequence of these findings is that moist compaction

should be used in the field to provide the greatest thermal conductivity in

case the backfill subsequently dries.


IV. BACKFILL TREATMENTS

Introduction

The most effective, controllable and promising ways to reduce or maintain

low thermal resistivity of well-graded backfill materials will be the develop­

ment of additives which can either (1) prevent water migration in the soil due

to temperature gradients, or (2) substitute for water in bridging the gaps be­

tween particles.

The following criteria were established as a basis for materials suitable

for successful backfill treatment in particular:

Inexpensive

Readily available and abundant

Non-toxic

Non-polluting

Stable and durable (permanence of treatment)

Easy to handle and mix

Effective with different soil types

Effective in small quantities

Treatment of Monterey No. 0 Sand

Several additives were tested first using Monterey No. 0 sand. Those

which indicated promise as regards their ability to reduce significantly the

thermal resistivity of this sand as well as some other subsequently obtained

materials were then tested using the Fire Valley thermal sand.

The results of the additive tests on Monterey No. 0 sand are summarized

in Table 2. The effectiveness of any treatment listed can be judged by com­

parison of the thermal resistivity with that for the untreated sand, which

for 0% water content is 340° C-cm/watt and for 10% water is 62° C-cm/watt.

4-1
TABLE 2

EFFECT OF ADDITIVES ON THERMAL RESISTIVITY OF MONTEREY #0 SAND

Average Dry Reduction


Additive P Water Density in p
Saaple (0C-cm/ Content kN/m 3 Notes
due to
No. watt) (%) (pcf) Additive

No edditive S-0421-1 16.2 Coflf>letely dried in 60a C oven for four


340 0
(103) days (Initial compaction at 10% water)

No Additive S-0421-1 16.2


62 10
(103)

5% cenent + 16.2 Dried in the oven after compaction at


201 0 40%
5% bentonite (103) 10% water

5% cenent + 16.2
48 10
5% bentonite (103)

5% NAjSO^ • 10 H20 ♦ 16.2


S-0512-1 70 10
5% cerbon black (103)

5% Na2S04* 10 H20 + 16.2


70 ? 80% Pried in 60° C oven for 7 days before
S% carbon black (103)
test

10% charcoal S-061S-1 15.3 8%


368 0
(97.2) increase

10% charcoal 15.3 Dried at 60® C for days


S-0615-1 310 0.5
(97.2)

15.3
10% calciua sulfate S-0623-2 91 10
(97.2)

15.3 Dried in oven 60® c for 4 days before


10% calciua sulfate S-0623-2 240 0.2 30%
(97.2) measurement

4% CaCl2 ♦ 15.3
S-0623-1 76 10
(97.2)
6% sodiusi silicate

4% CaCl2 + 15.3
S-0623-1 166 4.2 53% Dried at 60® C for 4 days
(97.2)
6% sodiia silicate

5% NaS04 • 10 H20 16.2


S-0629-1 64 10
(103)
5% charcoal '

5% Na2S04 • 10 H2o 16.2 Dried at 60® C for 4 days before i


S-0629-1 190 2.5 45%
(103) measurement 1
5% charcoal

16.2 2 week curing at room temperature


10% C-7 S-0713-1 68 10
(103) (68® C)

16.2
10% C-7 90 5 Dried at 60® C for 4 days
(103)

16.2
10% C-7 100 0
(103)
70% 14 day's drying in 60® c oven

16.5
1% copper fiber 145 0 57%
(105)

4-2
It may be seen that none of the systems tested yielded a material having

a thermal resistivity as low as that for the untreated sand at 10 percent

water content. On the other hand oven drying the untreated sand (while still

compacted in the mold) at 60° C for 4 days caused complete drying and nearly

a six-fold increase in p. Comparable drying of some of the treated specimens

did not cause nearly as great a loss of thermal conductivity.

The systems listed in Table 2 that are promising include:

1. 5% Cement +5% Bentonite

A 40 percent smaller value of p after drying than the value for dry un­

treated sand was obtained.

2. 5% Na SO. • 10 HO (Glauber's Salt) +5% Carbon Black


------z—4------- 2------------------------------------------

This combination of materials was the most effective (after oven drying)

of those tested. It has the major disadvantage, however, that the Na^SO^ •

10 H^O is water soluble. Thus in field application it would be washed out,

and contaminate the surrounding ground as well.

3. 10% Calcium Sulfate* 10

A small amount of water (0.2%) was retained after drying. This, in con­

junction with the enhanced interparticle solid contact probably account for

the 30 percent lower value of p.

4. 5% Na^,S0^ ‘ 10 H^O + 5% Charcoal

This system, while not as effective as the system containing carbon black

is still effective. It suffers the disadvantage of solubility of the Na2SO^

10 H20, however.

4-3
5. C-7

C-7 is a cement-based additive developed by Takanaka-Komuten, Inc., Japan,

as a stabilizing agent for soft, wet dredged materials. It appears to offer

good potential for maintenance of a low p, with a 70% smaller value than the

untreated sand after 14 days drying at 60° C.

6. Copper Fiber (1%)

Copper fiber added in small quantities resulted in a 57 percent decrease

in p for the dry sand. Economics and electrical conductivity considerations

argue against the use of this material, however.

The results in Table 2 pertain to samples after compaction, after curing,

or after drying. In some subsequent tests changes in p during curing and

during drying were measured. During curing (at 20° C or 60° C) the samples

were contained in molds and sealed at both ends to prevent water loss. During

drying one end was uncovered. Measurements of p were made daily, with equi­

libration of the sample at 20° C prior to each measurement. For samples

tested during a drying phase this resulted in a 16 hour drying period (60° C)

and an 8 hour equilibration period each day.

Fig. 31 shows the results of such a test on a C-7 treated sample of

Monterey No. 0 sand. It is important to note that the water contents and

water content changes shown are averages for the whole sample. Local vari­

ations within the sample are to be expected in view of the drying procedure

used.

Treatment of Fire Valley Thermal Sand

Materials that showed some promise for reduction of thermal resistivity

when mixed with the Monterey sand were tested using the Fire Valley sand,

along with some additional materials. The results of these tests are sum­

marized in Table 3.
4-4
S-0713-1
Monterey No.O+ C-7 (10%)
yd ' /5.2 kN/m3(103pcf)
Initial Water Content =10%
C -c m /W o tt)
Final Average Water ContentsO%
p(

Curing Period Drying Period


Average Water Content Loss (% )

Days After Compaction

FIG. 31 EFFECT OF C-7 ON THE THERMAL PROPERTIES OF MONTEREY NO. 0


SAND DURING CURING AND DRYING
It may be seen from the results in Table 3 that none of the additives

could effect a significant reduction in thermal resistivity compared to the

untreated soil in a moist (6% water content) state. This is not surprising

in view of the low value (~50 °C-cm/watt) for the untreated material when wet.

On the other hand several additives were effective in comparison to the un­

treated material after drying. A measure of this effectiveness is given in

Table 3 by the percent reduction in p due to additives.

Among the materials investigated that appear effective are:

1. Calcium chloride + sodium silicate—a combination which forms an in­

soluble silica gel.

2. C-7.

3. Chemicolime—a cement-based compound produced in Japan for soft

soil stabilization.

4. Portland cement.

5. Cement-bentonite mixture—it was hoped that this combination would

combine the effectiveness of Portland cement with the water-holding

capacity of bentonite. It was not as effective as cement alone.

The results show that of the organic materials investigated portland

cement and C-7 are the most effective. These materials offer the advantages

of fairly low cost. They introduce a disadvantage, however, in that they

harden the backfill thus rendering buried cables less accessible.

Effect of Polymers on Thermal Resistivity of Fire Valley Thermal Sand

Three water-holding polymers were tested to evaluate their effects on

the thermal resistivities of Fire Valley thermal sand: (1) General Mills

Water Absorber Polymer, (2) Dow Chemical Polymer, and (3) Union Carbide

Viterra. In each case the polymer was dissolved in water before mixing the

4-6
Table 3

Effect of Additives on Thermal Resistivity of Fire Valley Thermal Sand


CLimestone aggregate, quartz sand mixture)

Average Dry *Reduction Days Required


Sample P Water Density
Additives in p due to to Dry Sample Notes
No. (°C-cm/watt) Content kN/m3 Additives in 60 °C Oven
(%) (pcf)
19.6
No additive S-0921-1 49 6 7 Moist tamped
(125) .
No additive fl
97 0 19.6 7 Moist tamped
(125)
19.6
No additive S-0902-1 47 6 Moist vibrated
(125)
19.6
No additive M
90 0 Moist vibrated
(125)
No additive S-0722-1 135 0 19.6 Dry tamped/
(125) vibrated
M-179 (10%) S-0721-4 46 10 19.6 Moist tamped
fl25)
19.6
M 179 (10%) If
90 0 7% 19 Moist tamped
(125)
18.5
Super-Slurper (2%) S-0727-1 58 10 Moist tamped
(118)
18.5
Super Slurper (2%) If
104 0 -7% 14 Moist tamped
(118)
Thermally Conductive 19.6
S-1015-1 112 0 -15% Tamped
Epoxy (2%) (125)

CaCl2 (4%) S-0809-1 38 6 19.6


Sodium Silicate (6%) Moist tamped
(125)

CaCl, (4%) 19.6


S-0809-1 80 0 17% 18 Moist tamped
Sodium Silicate (6%) (125)

CaS04 (5%) 19.6


S-0819-1 44 6 Moist tamped
Carbon Black (5%) (125)
Table 3 (Cent.)

Average Dry
Density *Reduction Days Required**
Sample P Water
Additives kN/m3 in p due to to Dry Sample Notes
No. (°C-cm/watt) Content in 60°C Oven
(pcf) Additives
(%)
CaS04 (5%) 19.6
S-0819-1 99 0 -2% 12 Moist tamped
Carbon Black (5%) (125)

C-7 (10%) S-0807-2 48 6 19.6 Moist tamped


ri25'l

C-7 (10%) If
63 0 19.6 35% 16 Moist tamped
(125)
19.6
C-7 (5%) S-0831-1 48 6 Moist tamped
(125)
C-7 (5%) M
69 0 19.6 30% 11 Moist tamped
(125)
19.6 Moist tamped
C-7 (3%) S-0901-1 49 6
(125)
19.6 11 Moist tamped
C-7 (3%) M
81 0 (125) 16%
19.6 Moist tamped
Chemicolime (10%) S-0720-4 S3 6
(1251
Chemicolime (10%) ft
72 0 19.6 26% 10 Moist tamped
(1251
51 6 19.6 Moist tamped
Cement (10%) S-0907-1
(1251
Cement (10%) ff
78 0 19.6 20% 16 Moist tamped
(1251
Cement (5%) 19.6
S-0813-1 51 6 Moist tamped
Bentonite (5%) (125)

Cement (5%) 19.6


S-0813-1 85 0 (125) 12% 11 Moist tamped
Bentonite (5%)
Lime (5%) 19.6
S-0902-3 42 0 (125) Moist tamped
Bentonite (5%)

Lime (5%) 19.6


ff
101 6 -4% 13 Moist tamped
Bentonite (5%) (125)
Table 3 (Cent.)

Average Dry
*Reduction Days Required**
Sample P Water Density
Additives in p due to to Dry Sample Notes
No. (°C-cm/watt) Content kN/m3
(pcf) Additives in 60°C Oven
(%)

Chemistress (10%) 19.6


S-0827-1 58 6 Moist tamped
Gypsum (1%) (125)

Chemistress (10%) 19.6


M
86 0 11% 7 Moist tamped
Gypsum (1%) (125)

Embaco (5%) S-0902-2 44 6 19.6 Moist tamped


(125)
Embaco (5%) If
94 0 19.6 3% 11 Moist tamped
(125)

vo Reduction with respect to untreated sample dried to 0% water content after initial compaction by tamping at 6%
water content.

**0ne day = 16 hours drying and 8 hours cooling


water with the sand. Due to the swelling effect of the polymers, the maximum

dry unit weight that could be achieved for samples treated with either the

Dow Chemical Polymer or Viterra was only 18.8 kN/m3 (120 pcf), and that for

samples treated with G. M. Polymer was 19.6 kN/m3 (125 pcf).

The thermal resistivities were measured periodically while the treated

samples were dried in an oven at 60° C (140° F). Figs. 32, 33 and 34 show

the effect of drying on the thermal resistivities and water contents of these

samples. An untreated sample with 6.5% initial water content and 19.6 kN/m3

(125 pcf) dry unit weight also was prepared and subjected to the same drying

process for comparison. An attempt to compact an untreated sample with 10%

water content at 18.8 kN/m (120 pcf) dry unit weight failed, as the uncom­

pacted material had a unit weight greater than 20.4 kN/m3 (130 pcf).

Thermal resistivities of all dried treated samples range from 100° C-cm/

watt to 125° C-cm/watt, or about 10-25% higher than the untreated samples.

The high thermal resistivities of the dried polymer-treated samples probably

are due to the swelling effect of polymers which not only prevents the com­

paction to high density, but also separates the contacts between particles.

Thus, the results suggest that backfill treatments using these polymers would

be detrimental if complete drying were to occur.

On the other hand, it may be noted from Figs. 32 to 34 that the additives

delayed the rate of water evaporation from the treated samples. Thus exposure

to a high temperature for a given time period results in less increase in

thermal resistivity for the treated material than for the untreated material.

Accordingly in situations where high cable temperatures are periodic and some

replenishment of water in the backfill is possible in the intervals between

high temperatures, the use of these chemicals might be beneficial.

4-10
Thermal Resistivity (°C- c m /w a tt)
-Untreated Sand at a
Density of 18.8 kN/m3
(120pcf) -------- -

(136% ___ _________

1.0% Polymer

Untreated Sand at a
No te Polymer percentages-
■■

Density of 19.6 kN/m3 (125pcf) are by weight of


dry soil

1.0% Polymer

Untreated
Sand—
200
Time of Drying at 60°C (hours)

FIG. 32 EFFECT OF DOW CHEMICAL POLYMER ON THERMAL


RESISTIVITY AND WATER CONTENT OF FIRE VALLEY SAND
S 120
0.05%

—□—□---- ---
----------- 'J=l a~~0~5 % Polymer
Dry Unit Weight of AII Samples-19.6kN/m 3 _
Untreated Sand (125 pcf)
T h e rm a l

Note-- Polymer percentages


are by weight of dry soil—
W a te r C o n te n t (% )

0.5% Polymer

0.2%

UntreaU 0.05%
Sand -
200
Time of Drying at 60 °C (hours)

FIG. 33 EFFECT OF GENERAL MILLS POLYMER ON THERMAL


RESISTIVITY AND WATER CONTENT OF FIRE VALLEY SAND
-Untreated Sand at a Density of
18.8 kN/m3 . ____ a—
(120pcf)
V.2%

Untreated Sand at a
1.0 % Viterra Density of 19.6 kN/m^
(125 pcf)

Note: Viterra percentages


are by weight of dry soil -

tO% Viterra

0.5%

IOO 200
Time of Drying at 60°C (hours)

FIG. 34 EFFECT OF UNION CARBIDE VITERRA ON THERMAL


RESISTIVITY AND WATER CONTENT OF FIRE VALLEY SAND
To illustrate that the beneficial effects of the chemical treatments

depend mainly on water retention the data in Fig. 32 for the Dow Polymer and

Fig. 34 for the Viterra (sample density of 18.8 kN/m3 in each case) are re­

plotted in Fig. 35 in the form of thermal resistivity vs water content*. It

may be seen that the type or amount of chemical treatment had little effect;

although at the lower water contents the Dow Polymer treatment resulted in

somewhat higher thermal resistivities than did the Viterra treatment.

Waxes and Asphalt as Thermal Stabilizers

The use of wax as a backfill additive was investigated, and the results

are the most promising of any material yet studied. A shaping wax (CD 150/

160) obtained from Polygon Wax Manufacturing Company was mixed with dry Fire

Valley thermal sand at 71° C. The thermal resistivity values indicated in

Table 4 were obtained. Fig. 36 shows thermal resistivity as a function of

wax content. For wax contents of 2 percent or more by weight of dry soil a

thermal resistivity of about 50° C-cm/watt was obtained. This is essentially

the same value as obtained for this backfill when it contains 6 percent water.

Additives to the shaping wax did not lead to significant additional reductions

in thermal resistivity, as may be seen in Table 4. Results for a micro­

crystalline wax and paraffin are also given in Table 4. These materials

gave thermal resistivities comparable to that for the shaping wax.

As waxes are hydrocarbons, it was decided to investigate the potential

of backfill treatment using another type of hydrocarbon; namely, asphalt. It

may be seen in Table 4 that the thermal resistivity of asphalt-treated Fire

*As the densities of samples treated with the General Mills polymer are
higher than could be obtained with the other additives, data from
Fig. 33 are not comparable to that in Fig. 35.

4-14
140

o Viterra treated sample


• Dow Polymer treated sample

Dry Density^ 18.8 kN/m


(I20pcf)
C -c m /w o tt
(
Therm al R esistivity

Water Content (%)

FIG. 35 THERMAL RESISTIVITY AS A FUNCTION OF WATER CONTENT


FOR POLYMER-TREATED FIRE VALLEY THERMAL SAND

4-15
TABLE 4

EFFECT OF WAX ON THERMAL RESISTIVITY OF FIRE VALLEY THERMAL SAND

P Melting Point
Additive A Additive B
(0C-cm/watt) of Wax

3% Wax (CD 150/160) — 52 68° C

0
00
cn

o
6% Wax (CD 150/160) 6% Alumina 67

3% Silica
3% Wax (CD 150/160) Powder 45 68° C
#200-250

3% Wax (CD 150/160 3% Aluminum 49 68° C


Powder

3% Paraffin — 52 53.5° C

3% Micro-crystalline — 60 74° C

3% Asphalt (85/100) — 63 =88° C

Dry unit weight of all samples 19.6 kN/m3 (125 pcf)


(°C-cm/watt)

Soil dry density -19.6 kN/m


025pcf)
pResistivity,

y. - 20.4 kN/m &


(130 pcf)
Thermal

Wax Content (%)

FIG. 36 EFFECT OF WAX (CD 150/160) ON THERMAL


CONDUCTIVITY OF FIRE VALLEY THERMAL SAND

4-17
Valley thermal sand is comparable to that of wax-treated sand. Additional

studies using asphalt are in progress.

Because of the very favorable results obtained in the preliminary tests

using wax as an additive, more detailed studies were made.

Wax Selection and Properties. Some characteristics of waxes and considera­

tions pertinent to the selection of a material for backfill stabilization are:

(1) The thermal resistivity of wax is 320 thermal ohms, which is about

two times the thermal resistivity of water.

(2) The difference in melting points of different CD waxes are attributed

to their differences in molecular weight. The higher the molecular

weight, the higher the melting point.

(3) No melting or recrystallization is possible for CD wax at tempera­

tures under 30° C.

(4) Emulsified wax is available at a cost of about $1.70/gallon.

This material contains 50% emulsifying agent and 50% wax. >The

curing period of this wax is expected to be around 1 year under

normal temperatures.

(5) Slack wax, an unrefined wax, has similar thermal properties to CD

wax, but only costs 6<? per pound.

Preparation of Wax-Treated Samples. A mechanical mixer with variable tempera­

ture control was used to prepare the wax-sand mixtures. Known amounts of dry

sand and wax were poured into the mixing bowl and mixed continuously while

heating to 66° C (150° F). Then a weight of mixture to produce a one-inch

layer was poured into the standard mold with a thermal needle at its center.

A surcharge, having a hole at its center, was then placed on the sample and

an air vibrator was used to apply a dynamic load to the top of the surcharge.

4-18
The air pressure was supplied to the vibrator through a regulator which was

set at 0.5 kg/cm to 3 kg/cm according to the desired density. Vibrations

were applied until the height of the sample decreased to produce the desired

density. The surcharge was then removed and the procedures were repeated for

the successive layers. Four 25.4 mm (one-inch) layers and one layer 12 mm

(0.471 inches) thick were completed in this way.

The sample was then placed in a constant temperature room and allowed to

equilibrate at 20° C (68° F). Evaluations of the thermal resistivity of the

mixture were then made using the thermal needle method.

Effects of Waxes on Fire Valley Thermal Sand. Thermal resistivities measured

for Fire Valley thermal sand treated with CD wax and slack wax are shown in

Figs. 37 and 38. Fig. 37 shows thermal resistivity as a function of wax

content or water content. The performance of the CD wax treated sample is

very close to that of the sand containing water. The advantage of the wax

over water, of course, is that the wax does not migrate away or evaporate on

heating or drying of the samples, as does the water. As in Fig. 37, Fig. 38

shows that the slack wax is about 10%~20% less effective than the CD wax.

This might be attributed to the effects of unknown petroleum by-products in

the slack wax. Nonetheless slack wax proves to be an effective additive,

especially in view of its low cost. The maximum amounts of wax required for

Fire Valley thermal sand are 2% and 3% for CD wax and slack wax, respectively,

to attain the lowest thermal resistivity values.

Effect of Slack Wax on Monterey #0 Sand and Round-Robin Sand. Fig. 39 and

Fig. 40 show that slack wax also has the same effect as water on reducing the

thermal resistivities of Monterey #0 sand and Round-Robin sand. Round-Robin

sand treated with 6 percent slack wax produced the lowest stable value of

thermal resistivity (31° C-cm/watt) measured thus far in this project.

4-19
Sand and Slack Wax

Sand and Water

Sand and CO Wax

£ 60

o 40

Density of sand s 19.6 kN/m* 025 pcf)

Additive Concentration, Percent of Dry Soil Weight

FIG. 37 EFFECT OF SLACK WAX ON THERMAL


RESISTIVITY OF FIRE VALLEY THERMAL SAND

4-20
Dry Density (kN/m3)
20.5

No Additive
(moist vibroted,
then dry)

6% Slock Wax

6% CD 150/160 Wax

130
Density of Sand (pcf)

FIG. 38 EFFECT OF WAXES ON THERMAL RESISTIVITY


OF FIRE VALLEY THERMAL SAND

4-21
Soil dry density = 17.3 kN/m3 (NOpcf)

Flo 200

Sond and Wax

Sand and Water

% of Additive

FIG. 39 EFFECT OF SLACK WAX ON THERMAL


RESISTIVITY OF MONTEREY NO. 0 SAND

4-22
Soil dry density = 21.2 kN/m
(I35pcf)

Sand and Wax

Sand and Water

% of Additive

FIG. 40 EFFECT OF SLACK WAX ON THERMAL


RESISTIVITY OF ROUND-ROBIN SAND

4-23
Emulsified Wax Additives. Emulsified waxes are readily available and can be

mixed directly with backfill materials without the necessity for heating the

soil-wax mixture. An emulsified wax containing 50 percent wax and 50 percent

emulsifying agent was mixed with Fire Valley thermal sand. Samples were com­

pacted to a dry unit weight of 18.8 kN/m3 (120 pcf). After an initial measure

ment of thermal resistivity, samples were placed in a 60° C (140° F) oven to

accelerate drying. Values of sample weight and thermal resistivity were

measured periodically until constant values were obtained. The results are

shown in Fig. 41.

It may be seen that about the same reduction in thermal resistivity can

be achieved using emulsified wax as can be attained using slack wax. About

twice as much emulsified wax is required, however. This would be expected,

because the wax content of the emulsion is only 50 percent.

Effect of Temperature and Temperature History on Wax-treated Sands. The ef­

fect of temperature and temperature history on wax-treated samples has been

investigated. Samples of slack wax-treated Round-Robin sand and Fire Valley

thermal sand were prepared and then sealed with thermal foam plates. Sam­

ples were then put in a constant-temperature chamber; measurements of the

thermal resistivity were done after samples achieved different stable tem­

peratures.

Figs. 42 and 43 show the effect of temperature and temperature history

on the thermal resistivities of wax-treated Fire Valley thermal sand and

Round-Rdbin sand, respectively. The temperature of the chamber in these

tests was increased from 5° C to 80° C, and then samples were cooled back

to 5° C.

From the results, it can be seen that the change of thermal resistivities

measured at temperatures below the melting range is minimal. However, thermal

4-24
O Emulsified Wax treated sample
• Slock Wax treated sample ~

Dry Density -18.8 kN/m3


(120 pcf)
Therm al Resistivity (°C -cm /w att)

Original Wax Content (percent of dry soil wt.)

FIG. 41 EFFECT OF EMULSIFIED WAX ON THERMAL RESISTIVITY


OF FIRE VALLEY THERMAL SAND
20 40 60 80
Ambient Temperature (°C)
FIG. 42 EFFECT OF TEMPERATURE ON THERMAL RESISTIVITY
OF SLACK WAX-TREATED FIRE VALLEY THERMAL SAND

4-26
too T T
O Heating
A Cooling
Dry Unit Weight=2!2 kN/m3
80
Thermal Resistivity (°C-cm/watt)

(135 pcf)
Wax Content =6%

0L _|_______ |_______ 1_______


O 20 40 60 80
Ambient Temperature (°C)

FIG. 43 EFFECT OF TEMPERATURE ON THERMAL RESISTIVITY


OF SLACK WAX-TREATED ROUND-ROBIN SAND
resistivities measured at ambient (chamber) temperatures higher than the

melting point of the wax show an increase of 10-15%. No temperature hysteresis

effects on the thermal resistivity were observed in these two series of tests.

Evaluation of 3M Stabilizer

A 3M thermal stabilizer* has been tested and compared with C-7 and wax.

Table 5 summarizes the results obtained for treated Fire Valley thermal sand

and Monterey #0 sand. The results show the 3M stabilizer is about as effec­

tive as C-7. Both additives reduce the resistivity of the untreated Monterey

#0 sand by a very significant amount (about 60%). The improvement in thermal

resistivity of Fire Valley thermal sand treated with these two additives is

about 20%. In all sands studied, wax was the most effective material for re­

ducing and maintaining a low thermal resistivity, however.

Conclusions

No additive has been found which is more effective than water in imparting

a low thermal resistivity to a backfill material. On the other hand water

migrates and evaporates from hot backfills. Three water-holding polymers

have been found effective in retarding the loss of water, thus enabling the

treated backfill to retain a lower thermal resistivity for a longer time.

If dried completely, however, the thermal resistivity may be higher than if

no polymer were used. This is primarily because the water-holding polymers

make compaction to high density difficult.

Admixtures which substitute for water but which do not evaporate or

migrate away from hot spots have been investigated. Among the many additives

tested, waxes appear the best suited in terms of effectiveness in reducing

thermal resistivity, ease of incorporation and handling, and economics.

*3M Soil Thermal Stabilizer Lot #601501

4-28
TABLE 5

THERMAL CONDUCTIVITY VALUES OF SEVERAL BACKFILL-ADDITIVE SYSTEMS

!
Dry
Water Content Density
i
Weight % at Test kN/m3 P
Soil Type Additive of Additive % (pcf) 0 C-cm/watt

Ottawa Sand* None — 0 15.7 (100) 273


II II* 3M Not stated 0 16.5 (105) 98
Monterey #0 Sand None — 10 16.2 (103) 62
II II 11 It
— 0 16.2 (103) 340
i

II 11 11
3M 5 10 16.2 (103) 87
6Z-

II II 11
3M 5 0 16.2 (103) 130
II II II
C-7 5 10 16.2 (103) 68
It II
C-7 5 0 16.2 (103) 100
"
II 11
Wax CD 150/160 6 0 16.8 (107) 82
"
1 » II 11
Wax CD 150/160 2 0 16.2 (103) 120
i

Fire Valley Thermal Sand None — 6 19.6 (125) 47


II II It II
None — 0 19.6 (125) 90
II II 11 II
3M 6 6 19.6 (125) 39
II II II 11
3M 0 0 19.6 (125) 72
II II II
C-7 5 6 19.6 (125) 69

» 11 M 11
C-7 5 0 19.6 (125) 52

*Reported by 3M Company
Preliminary tests indicate that waxes can produce a permanent reduction in

thermal resistivity, will be non-polluting, and produce a treated soil which

will still allow easy access to the buried cable. This would not be the case

with other additives such as Portland cement and C-7 which, while effective

in reducing thermal resistivity, produce a hardened mixture.

Further study of the details of wax treatment and the durability and

thermal stability of treated materials is planned, as is field testing.

Because of the great promise shown by wax-treated systems, patent protection

is being applied for.

4-30
V. TEMPERATURE DISTRIBUTIONS AROUND BURIED
CABLES AND HEAT FLOW ANALYSES

Introduction

Evaluation of the transient and steady-state temperature distributions

and heat flows around buried power cables is useful for several reasons.

They include:

1. Determination of the maximum temperatures to be associated with

different cable loadings.

2. Study of the effects of different trench geometries.

3. Evaluation of the importance of the different backfill and natural

ground properties on temperature distributions and heat flow.

4. Determination of the time required for establishment of thermal

equilibrium of the temperature changes associated with cable

loading of given duration.

5. Evaluation of the effects of weather and solar radiation on tem­

perature distributions.

In this section previous methods for analysis of temperature distribu­

tions are summarized, a general analysis procedure for heat flows and tem­

perature distributions is presented, and examples of solutions for several

cases are presented.

The problem to be analyzed is the distribution of temperature around a

buried electrical cable as shown below. Both transient and steady state

conditions are of interest.

5-1
where T1 is the temperature of the cable surface

To is the initial soil temperature

kl, ko are the thermal conductivities of the trench backfill

material and the surrounding soil, respectively.

Generally k, > k , T, > T . The ambient air is assumed to be an infinite


1 o 1 o
heat sink in some cases, although in most instances reasonable accuracy in

prediction will require consideration of ground surface temperature variation

and solar radiation.

Previous Solutions

An early steady state solution, known as the Kennelly formula (Bauer

and Nease, 1958) assumes that temperature gradients in the soil are analogous

to the voltage gradient in an electrostatic field between two energized para­

llel conductors. The Kennelly formula is:

AT = 0.00522 Wpln (8 )
P

where W heat loss per foot of cable (watts/ft)

P soil thermal resistivity (°C-cm/watt)

At change in temperature at point p


P
d' distance from image of heat source to 'p' (cm)

d distance from center of cable to p (cm)

5-2
CabU Image

£>uried Cable

This method allows only a single constant value of soil resistivity for

the entire soil environment surrounding the cable.

In a method proposed by Schmill (1960) the cable soil environment is al­

lowed 2 differing values of conductivity; one for the ambient earth and the

other for the cable trench backfill: These values are held constant for the

analysis, which yields a steady state solution for the temperature distribu­

tion.

The thermal potential from a source of q units per unit time per unit

length, at a distance of r units, through a medium with conductivity c^, is

given by:

<j> = - — in r2 (9)
c.
where k is a constant.

If a second medium is introduced and separated from region (1) by boundary

'A', two conditions must be satisfied.

1) the potential at any point of the


A
boundary should be continuous.
C
'■"I C
'-2 2) the normal component of the heat

flux density should be constant at


o
both sides of the boundary, at any

point on the boundary.

5-3
To satisfy these boundary conditions two additional ficticious line sources

and must be added.

ft

q^ is a heat source of intensity: q^ = mq

q^ is a heat sink of intensity: q^ = nq

and is coincident with q.

m and n are constants

The potential at point p^ is:

. q ,
- k — In r
2 - k "2 _
— In r
2
(10)
ci 1 °i 2

The potential at point p2 is:

- k -2. in r? (ID
C2 1

Applying appropriate boundary conditions yields:

C1 ~ C2
q and q , - [i^l
C1 + C2 1 L ci * “oJ

as q^ = mq, this implies

m = (12)

Thus m is a coefficient which takes into account the different conducti­

vities and transforms both regions to a single uniform medium.

Now allowing the geometry of the boundary between the two layers to

change, as shown below.

5-4
G
%

q2 = hkj

q3 = mq

q4 = -mq

q1 = (l-m)q
A

Finally allowing thr boimdary geometry to take the shape as shown below,

requires additional ficticious heat sources.

By relating q to q^ as above and

applying the boundary conditions


0%
a complex expression for the
Q
potential at any point is obtained.
%o V o%

To simplify the expression, consider only those points where a potential

plane cuts the axis drawn vertically through the cable center. The expres­

sion then becomes:

k — |m ( ) + m2 ( ) + m3 ( )

2 \2
(2h-R+2b2) 4a^+ (2h-R) 2 4a2+ (2h-R) 2
, kq \ (2h-R)2
((), = — ; In ---- — + m x ----- -—~— X *
1 C1 R2 . 2 2
(2b2+R) 2
K 4ai+R2 44*rZ

4aJ+(2b2+R)2 4a2+(2b2+R)2
x —----------- x —----------- (2h-R+2b)2\2 /4a2+(2h-R)2X2
+ m
4a1+(2h-R+2b2)2 4a2+(2h-R+2b2)2 (2b+R)' 4a2+R2

4a2 + (2b2+R)2 N2 4a^+(2h-R+2b2)2 4a2+(2h-R+2b2)2 4a^+(2b+R)2

V4a2 + (2h-R+2b2) 2 4a^+(2b2+R) 2 4a2+(2b2+R)2 4a^+(2h-R+2b)2

5-5
4a2+(2b+R)2 't2h-R+2(b+b2)]2\2 4(a+a1)2 + (2h-R)2 4(a+a2)2 + (^^

4a2+(2h-R+2b)2 ] + "3 [’ [2(b+b2)+R]' 4(a+a1)2+R2 4(a+a2)2+R2

^4a2+(2h-R+2b2)2Xz 4ai+(2h-R+2b) 2 4a2+(2h-R+2b) 2 4 (a+a^^) 2+(2b2+R) 2


x ---- ---------- x ---- ---------- x --------------------
2.,„, . 2 . /oi. . T,\ 2 A,-.- s 2 , fyu x 2
4a2+(2b2+R)2 4a1+(2b+R) 4a2+(2b+R) 4(a+a1)^+(2h-R+2b2)

4(a+a2)2+(2b2+R)2 4a2+ (2b+R) 2 4aJ+[2(b+b2)+R]2 4a2+[2(b+b2)+R]z

4(a+a2)2+(2h-R+2b2)2 4aZ+(2h-R+2b)2 4a^+[2h-R+2(b+b2)]2 4a2+[2h-R+2(b+b2)]

+ m** ( ) .

where R is the distance along the axis from the cable center to where the c|>^

equipotential plane intersects the vertical axis. The series converges rapidly
a
so the m term is considered to give sufficient accuracy. The geometric con­

stants a^, a2» a, b^, b2, b, h are defined as shown.

£ a.r^ Sur P (XCc

ca
_r , ■
c, J8
....-< )
0.2.
a

According to Schmill (1960), the solution can easily be transformed to

calculate the shape of the complete equipotential surface point by point.

This method allows for regions of 2 differing conductivities in the

correct geometric spaces, but the conductivities must be assumed constant

with temperature.

Slaninka (1974) has developed a solution which allows the thermal con­

ductivity to vary with temperature- The conductivity is assumed to vary thusly:

k = ko (1 + aTn) (14)

where ko, a, n are constants which are valid for the given water content of

the soil.
5-6
The governing differential equation for steady state heat flow through

a solid is: (r is the symmetric cylindrical variable)

8 T
r\ 2
1 8t
r 8r e (i(S) (15)

Inserting the derivative:

8k _ 8k j)T 8k_ 8w
(16)
8r 8t 8r 8w 8r

8k
and allowing = 0 [implies non-linearity]
8t

8k
0 [implies homogeneity] where w = water content

yields:
82t ^ 8t 1 8k /8t\2
(17)
r 8r k 8t \8ry

for which the solution is:

F(T) = c. In — + c^
1 r 2

T
where F(T)
/
J T„
1
kdT (18)

Solving these equations using complex variables gives a relation between

F(T) and cp (potential) .

The tracks of equipotential planes are stated to be circles with:

£a.rik. Surface
radii T
cj)2 - i
Buried
and
(j)2 + 1
centers = -H ,2
<|) - 1

5-7
The equivalent thermal conductivity is given by:
_n+l _n+l
, [1 + _5_ Tl T° 1 (19)
V° L n + 1 Tl " To J

the external thermal resistance of the cable is given by:

11/2-
(20)
Sir -[f •[(?)’->] ]

This is similar to the Kennelly formula; i.e., p In


[■ [va]]
It is then supposed that the soil has two different thermal conductivities,

possibly due to moisture migration:

Ear+k SurS-dce.

- ^ T -N

First the temperature of the equipotential plane that separates the

regions of different k must be found. (T^)

This is accomplished using a trial and error process.

First assume T„ T + — (T - T )
0 2 K 1 O'
_n+l _n+l
and calculate (21)
L "+1 h - t2 J
r mn+l mn+l 1
k = k (22)
e2 0 L n + 1 To - T J

and
(2H - e ) e2
(23)
2TTk e (2H - e2)
]
2H - e.
(24)
2TTk
'2 J
5-8
T2 is calculated from the relation:

T + ---------- (Ti - T0> (25)


0 He, + He,

If T2 calculated does not equal T2 assumed, another assumed and

the process is repeated until T„ assumed = T„ calculated. All He. referred


2 2 3

to or used from this point on are the final He_. calculated using the final

T2 above.

In the first zone (k^) : the equipotential planes are circles with:

center = -H -
2(H - e)
2He,
where e
(j)1(2H - e2) + e2

e(2H - e)
radii
2(H - e)

In ,the second zone (k2) : the equipotential planes are circular

sections with:

e(2H - e)
radii
2 (H - e)

where e

centers = -H -
2(H - e)

The solution gives a steady state distribution of temperature for the

case of temperature dependent soil thermal resistivities. Two regions of

differing conductivity may exist.

Schmill (1967) presents a steady state analysis for the case of variable

soil thermal resistivity. This analysis uses the Kennelly formula to deter­

mine temperature distributions. Formulas are introduced to calculate the

eccentricity of the isothermal circles resulting from the infinite heat sink

at the ground surface. A method is presented to account for moisture migration

in determining an overall effective soil thermal resistivity.

5-9
A hypothetical pattern of moisture content in a uniform soil is assumed:

The values of thermal resistivity in

the r, - r. region and the r_ - r.


12 3 4
region are labeled p and p
M m
respectively.

The resistivity at some r where

r2 SrSr3 can ke foun<* by finding

w = w + --- -- (r - r ) (26)
X r3 ” r2

and substituting into

PM 10_SlWr + p 10_S2Wr (27)


M m
E>'>sW1ce. from HcoJ- Source

Equation (27) is an empirical equation, where s^ and s2 are moisture

migration coefficients.

Using Fourier’s equations for steady state heat flow.

<*T\= o (28)
dr \ p dr y

If the constant heat rate is q watts per unit length of the cable, then:

a dT
(29)
p dr

in which a is the area.

Combining (28) and (29) yields:

_ -2_ £ (30)
dr 2tt r

which is the differential form of the Kennelly formula.

Integrating (30) from r^ to r^, that is:

5-10
•/m
I
4
da n-
Jr4
^ £ dr
2tt r

f*! - 3- £. _ -3L
T, - T,
A 27T r f
2TT Jta ~r (31)

Since p is not constant from r„ to r, , but is constant from r, to r„ and


4 1 12
from r^ to r^, equation (31) can be rewritten as

T - T (32)
1 4 2tt
[/r*f-*/r2f-*/ri^'
l/r4 ^2

from which: T, - T. =: o P In !i + p £ dr + p In — (33)


1 4 2 tt m * 4* •'r
J r M r
1J

T - T _q_ r3 r2
1 *4 P In — + I + P,, In — (34)
2ir m x„ kM r,
4 1
r2+r3
To evaluate I in the most simple form, take r^ = —-— and substitute into

(26) and (27) to determine p .

r r r
Then T, - T. = - fpM ln — + PT in — + p In —) (35)
1 4 2tt V M r2 r Km r )
2 * '3 ^4

An eccentricity factor 'a1 is introduced to account for the heat sink at

b + a = 2h (36)
c 2
b = C1 (37)
a
2 2 V2
h - [h2 - C^] (38)
5-11
It is assumed that the constant moisture content contours develop

the same eccentricity. That is, p from c to is still PM at all points.

The distances r^, r^ and r^ define this more practical situation (with r^

cable radius): i

^r
CO (si
(U V 02
Ground Surface
<

c '

e
tr
* ■

°L

c^ ^ r^, ^2 ^ r2 ^ecause °f eccentricity.

The resistance to heat flow from r^ to r^ using Kennelly's formula is:

% 2tt (ln r;L ln r2)

_ _ N
(39)
RM 2tt \ln r, R_ ,

Similarly:

II (40)
2tt
(*- ? • £

But r. r4, so

pm / R
(41)
2?^

In order to eventually develop an expression for p a single value of


ef f
p to simulate the combined effect of the different resistivities present,

establish the relation:

R, R.
T - T T - T
1 4 1 GROUND ■J'VW = 3? Oeff [ln ^ - ln (42,
SURFACE

5-12
By substituting (39), (40) and (41) into (42) and simplifying:

R1 r2 R2 r3 R3
P„ In — — + pT In — + p In — (43)
M rl R2 1 r2 R3 m r3

Therefore the temperature rise of any point in the soil mass above the

soil ambient temperature is:

at . A p 1„ £ (44)
2tt eff r

where r is the distance from the cable center to the point and R is the

distance from the point to the cable image.

Discussion of Solutions

The main advantage of the Bauer and Nease (1958) method is the ease of

computation. The only parameters required are W, the heat loss per linear

foot of cable, and p, the soil thermal resistivity. This solution, because

of its simplifying assumptions, is not very representative of the actual con­

ditions. The presence of a trench with a backfill of different thermal

properties cannot be included.

By requiring a single constant value for the soil thermal resistivity,

the variation of thermal resistivity with temperature and moisture constant

must be ignored. The soil is also assumed to behave isotropically with

respect to thermal resistivity. This solution deals only with steady state

conditions.

Calculations and a graph are presented by Bauer and Nease (1958) which

show excellent agreement between the calculated temperature distribution and

the measured distribution. However, the value of soil thermal resistivity

was not measured, but rather it was backcalculated from the Kennelly formula

and known temperature vs. distance values. This implies that for the given

5-13
assumptions and an accurate value of thermal resistivity, this solution does

well in predicting the temperature distribution.

The advantage of the analysis presented by Schmill (1960) is that it con­

siders a rectangular trench with a different conductivity than the surrounding

earth. It assumes both the trench material and the surrounding earth to ex­

hibit isotropic conduction properties. The analysis does not consider vari­

able conductivities, dependent on temperature or moisture content, and it is

applicable to steady state conditions only. The final solution is complicated

and may require a computer for efficient evaluation. No comparison is made

between a temperature distribution as predicted by this analysis and a mea­

sured distribution.

The Slaninka (1974) analysis accounts for variable soil thermal conduc­

tivity with temp'erature for a given water content. It also allows that at a

given distance from the heat source the water content may change, and hence

the conductivity may change. This feature can be used to account for mois­

ture migration away from the source. The geometry of the plane where the

conductivities change is limited to circular and hence only an approximation

of a rectangular trench can be made. All soils are assumed isotropic with

respect to heat conduction properties. The function that relates conducti­

vity to temperature at a given water content may be difficult to obtain with

accuracy. The analysis for water migration involves an iterative solution

which may require a computer to be practical.

The analysis is for steady state conditions only. No comparison is

made between a temperature distribution as predicted by this analysis and a

measured distribution.

The advantages of the Schmill (1967) analysis include the ability to

handle a variation of moisture content, and hence a variation in resistivity.

5-14
with distance from the source. Part of the analysis deals with the effect

of the infinite heat sink at the ground surface on the temperature distribu­

tion. The input parameters required include the variation of water content

with distance from the source; the migration coefficients which relate the

resistance of an intermediate point with the resistances and water contents

in neighboring regions; and the depth and intensity of the source. The lines

of equal water content are assumed circular about the source and therefore

only an approximation of a trench can be made. The migration coefficients

may be difficult to obtain for practical use. All soils are assumed to have

isotropic heat conduction properties. The analysis is applicable to steady

state conditions only. No comparison is made between predicted and measured

temperature distributions.

From this review it was concluded that none of the previously available

solutions for steady-state temperature distribution around buried electrical

cables accurately models the true geometric conditions or properties. None

allows for evaluation of transient conditions; thus times for attainment of

equilibrium or the influences of cyclic thermal loadings cannot be predicted.

Program HEAT

The finite element computer program HEAT, developed by Professor R. L.

Taylor of the University of California, Berkeley (Taylor, 1975), has been

found to be well-suited for the analysis of both transient and steady-state tem­

perature distributions both within the cable trench backfill and the surrounding

soil. This program can determine heat flows in systems of irregular geometries

and complicated boundary and initial conditions, which cannot be analyzed using

analytical closed form solutions. The basis of this program and its applica­

tion are outlined below.

5-15
The method is based on the governing heat flow equations:

Vq + pcT - Q = 0 (45)

which is the heat balance equation; and Fouriers Law

q = -JcVt (46)

which expresses heat flow rate in terms of thermal conductivity and

temperature gradient. In these equations:

q = heat flow rate

T = temperature

Q = heat generation rate

c =specific heat

k = thermal conductivity

p = mass density

V = gradient operator. Combining (45) and (46)

yields: - V(kVT) + pcT - Q = 0 (47)

Using the calculus of variations on equation (47) yields:

D(T) + AT = B (48)

where D, A and B are constants.

If the continuum earth is divided into a finite number of rectangular elements

with nodes at the intersection of element boundaries, then equations (45) and

(46) are valid for each element. Therefore equation (48) is valid for each

element, and if the temperatures at all points within an element are assumed

to be linear functions of the temperatures at the boundary nodes, the resul­

tant equations are a set of linear equations in T and Q. By applying the

boundary conditions and the initial conditions, equation (48) can be solved

on the computer to yield the temperatures at the nodes at time t.

5-16
Since equation (48) is equally applicable to all elements, the input

parameters required to calculate D, A and B can be specified for each parti­

cular element. Anisotropic conductivities for each element may be entered.

The continuum discretization into elements may take numerous shapes, and

therefore the presence of a backfill trench can easily be taken into account.

Various initial and boundary conditions with respect to temperature and heat

flow may be entered. A transient analysis as well as a steady state analysis

is possible. Other parameters required for analysis include the mass densi­

ties and the specific heats.

This analysis requires a computer; however, this method imposes the

least limiting assumptions, has a large amount of user options, and most

realistically analyzes the problem.

To test this program a comparison was made between the values of tempera­

ture vs. distance predicted by the program for a thermal needle experiment and

the values computed using the analytical solution (Carlslaw and Jaeger, 1959).

The results, shown in Fig. 14, indicate excellent agreement.

The finite element solution was used also to calculate temperature in­

crease as a function of time at the center of a sample during a hypothetical

thermal needle test. From the values obtained a value of p was calculated

that agreed with value of p initially assumed. This result provided further

evidence of the usefulness of program HEAT for study of heat flows in this

investigation.

Steady State Heat Flow Analyses

Heat flow analyses of typical cable-trench arrangements using computer

program HEAT have been performed. The first typical cable-trench geometry

chosen for analysis is shown below.

5-17
Ground Surface

p varies

from trench
.91 j m

0.61 m (2')
I---------- -

The first step in applying program HEAT to this problem was to discretize the

semi-infinite soil into a finite cross sectional area suitable for analysis.

Since the problem is symmetric with respect to a vertical line drawn through

the cable center, only the right or left symmetric section need be analyzed.

The factors which controlled the size selected were the boundary conditions

that were to be specified around the perimeter of the smaller area. After

studying the available options in program HEAT it was determined that the

most suitable method of specifying these boundary conditions was to use a con­

stant temperature boundary.

The next step in the analysis was to vary the size of the discretized

area so that the computed temperature distribution would not be significantly

different from that obtained for a free field. After performing several tem­

perature distribution analyses, it was determined that a 6.1 m (20 ft) wide

by 6.1 m (20 ft) deep discretized area was the minimum for dependable results.

It was then assumed that for this area, a 51 mm (2 inch) diameter cable could

be considered as a line source. The 6.1 m by 6.1 m area was then subdivided

into a group of smaller finite elements. The typical finite element mesh lay­

out used is shown in Fig. 44.

5-18
0.305mW)
M
l* St t* T» 94 04 HS Its 137 149 159 170 191

1.22m
!6 26 42 j 52 68 78
rTrsnch . tid9
9% 7 /IS 128 138 148 158 168
*549 r5
!4
a 39 49 39 T9 9t
\
y
94

< (104
—Cc >/0
it 99 npO\ ttJ
12 pt22 ST y
it qo/
4.27 / to UP

x m
8 /8 34 44 60 70 86
2
fi y84\ ttO
-n 305

5 3/ 57 83
s
3

4
f.aJ

3
1.22

0.6/ l
ti

J 27 79 909 999 929 939 949 959


S

97 tos_____ M3_____ 'SL /J9


t*7_____ JX_____ t49 !S2.
/so______ >7
I

n !/
O 97 S3 79 ___ 105 ISZ_____
i 0.6! 1.22 183 2 44 3.05 3.66 4.27 4.88 5.49 6.10
'—Origin 0.6im(2'i

FIG. 44 FINITE ELEMENT MESH USED FOR ANALYSIS OF TEMPERATURE


DISTRIBUTIONS AROUND BURIED POWER CABLE

5-19
As a check on the validity of the assumptions made in setting up the

finite element analysis, a temperature distribution analysis using the finite

element method was made for a homogeneous soil system; all soil around the

cable had the same thermal resistivity. The temperature distribution for

this same soil system was also calculated analytically using the Kennelly

formula. The results of these two analyses were compared and found to yield

the same temperatures to within 5% at a distance of 2.74 m (9 feet) from the

cable. Therefore, it was concluded that the finite element analysis yielded

satisfactory results and could be used to predict temperature distributions

for more complicated systems.

Allowable Heat Input - Homogeneous System. The variation of allowable heat

input rate to the soil to limit the cable sheath surface temperature to 90° C

as the resistivity of the homogeneous soil system is varied and is shown in

Fig. 45. As would be expected, the lower the soil resistivity, the higher

the allowable heat input ratio.

Allowable Heat Input-Trench System. A trench of variable thermal resistivity

was next considered. The allowable heat input rate to the soil to limit the

cable sheath surface temperature to 90° C was calculated for surrounding

ground with p = 150° C-cm/watt. A diagram of this system and the results of

the analysis are shown in Fig. 46. Also plotted on this figure are the re­

sults for a homogeneous soil system of p = 150 and a homogeneous soil of p

equal to the backfill resistivity.

The gain in allowable heat input by installing a lower resistance back­

fill is the difference between the middle and lower curves. If the trench

size was extended to infinity in both directions, the gain in allowable heat

input would be the difference between the upper and lower curves. By comparin

5-20
C able S h e a th S u rface Tem perature to 9 0 °C (w otts/ft)
T=20 C

0.915 m Initial T = 20 C
13') p - variable
A llo w a b le H e a t In p u t R ote to the S o il

Watts /m
Resistivity of Soil (thermal ohms)

FIG. 45 MAXIMUM HEAT INPUT RATE TO LIMIT CABLE SHEATH


TEMPERATURE TO 90°C IN A HOMOGENEOUS SOIL SYSTEM
9 0 ° C (watts/foot)

p=varies
Uowobte H eat Input Rote for Coble Surface Temperature *

0.61m

Wotts/m
Homogeneous p - trench resistivity,
ie. infinite size trench

!50/i r

Homogeneous p =I50/l

Resistivity of Backfill Soil (thermal ohms)

FIG. 46 ALLOWABLE HEAT INPUT RATE TO LIMIT THE CABLE


SHEATH TEMPERATURE TO 90°C FOR DIFFERENT
TRENCH BACKFILL RESISTIVITIES

5-22
these 3 curves it is seen that a significant increase in allowable heat input

rate is achieved with a trench of the dimensions shown.

This figure can also be used to calculate equivalent homogeneous resisti­

vities for a trench/natural soil system. For example, consider the trench

system with a backfill p = 70 B and natural soil p = 150 B. By moving to the

left at constant allowable heat input rate until the upper homogeneous curve

is reached, (follow the dotted line) and then reading down, the equivalent

homogeneous resistivity for this system is found to be 90 B.

Effect of Trench Size and Configuration. The effect of trench size and config­

uration on the allowable heat is shown in Table 6. These results show that a

significant increase in trench size did not result in significant increases

in allowable heat input rate. Therefore, for economic reasons, a small trench

as is ,practically possible seems to be the logical choice.

If, as is common practice, the top of trench is backfilled with the

natural ground instead of a special backfill, the behavior shown in Fig. 47

is obtained. The allowable heat input rate was only decreased a modest

amount, even for a 0.915 m (3 ft) thick layer of native soil.

The effect of resistivity of the surrounding soil was also analyzed. A

diagram of this system and a plot of the results are shown in Fig. 48. These

results indicate that the allowable heat input rate is much more sensitive to

a small change in the backfill resistivity than a large change in the sur­

rounding soil resistivity.

Effect of Finite Cable Size. Fig. 49 compares, for the conditions shown, the

allowable heat input rates for a line heat source, as assumed for the above

analyses, and a cable of finite cross section. A 203 mm by 203 mm (8 inch by

8 inch) square cross section has been chosen instead of the more usual cir-

5-23
TABLE 6

ALLOWABLE HEAT INPUT RATE FOR DIFFERENT TRENCH CONFIGURATIONS

1.22m X 0.61m 1.525m X 0.61m 1.22m X 1.22m 0.915m X 0.61m 1.22m X 0.61m 1.83m X 0.61m 1.83m X 0.61m
Soil
d = 0.915m d = 0.915m d = 0.915m d = 0.915m d = 0.61m d = 1.525m d = 1.525m
Resistivities
/41 X 2’\ / 5 ' X 2 '\ /4 ' X 4'\ /3 1 X 2'\ / 4 ' X 2'\ / 6 ' X 2'\ /6 1 X 2'\
(trench/ambient)
U = 3' / \d = 3' j Vd = 3' / Vd = 31 / \d = 2' / \d = 5' / Vd = 5' /
.915m (3') top
cover

ALLOWABLE HEAT INPUT (WATTS/M)


PZ-

100/150 98.16 99.02 102.95 89.84 112.79 85.25 83.28

90/150 105.25 106.23

80/150 113.64 114.75 121.97 100.98 133.11 97.05 92.46

70/150 124.0 125.90

60/150 137.18 139.67 151.48 118.69 165.25 114.10 104.76

50/150 154.69 157.38

40/150 179.61 183.61 205.90 152.79 224.92 142.95 119.34

30/150 219.15 223.61


Allowable Heat Input Rate for Cable Surface Temp.=90°C (watts/foot)

1.83m 1.525m
p= varies

Entire trench treated for


tower resistivity

a
Wts
t/m
Top 0.9/5m 13'j of trench
natural ground at p-tSOA. _

Resistivity of Backfill Soil (thermal ohms J

FIG. 47 EFFECT OF NATURAL GROUND INSTEAD OF SPECIAL BACKFILL


IN UPPER PORTION OF TRENCH ON ALLOWABLE HEAT INPUT
250

ui 200
-26

Watts/m
150

-too

Resistivity of Backfill Material (thermal ohms)

FIG. 48 INFLUENCE OF NATURAL GROUND RESISTIVITY ON THE


ALLOWABLE HEAT INPUT TO A CABLE TRENCH SYSTEM
Ta - 20°C, No radiation

0.9/5m
(30
° W
C(t
cs
tf
/)
t

Cable as a line heat source


0
-9

O
O 203mm
square r203mm
cable (8"*8")
leSr
ufc
a e
eTp
m.

m
b
a

/
C

s
t
r

t
o
ef

a
W
l
Alw
oal
b e
eHatn
IptR
u t
a

’ 80 60 40 20
Resistivity of Backfill Soil (Thermal Ohms)
FIG. 49 COMPARISON BETWEEN LINE HEAT SOURCE AND CABLE OF
FINITE CROSS SECTION ON ALLOWABLE HEAT INPUT RATE
TO LIMIT CABLE SHEATH TEMPERATURE TO 90°C

5-27
cular cross section to facilitate computations by the finite element program.

It may be seen that assigning a finite cross sectional area to the cable leads

to a significant increase in the allowable heat input. Thus the assumption of

a line heat source is conservative.

Parameter Study* 50

The large number of variables influencing the rate of heat flow and tem­

perature distributions around buried cables precludes the development of gen­

eralized charts and graphs applicable to all cases. Any particular case can

be studied in detail using finite element program HEAT, however.

In the analyses reported thus far a heat sink at 20° C has been assumed

for the air above ground surface. In reality the ground surface represents

a convection boundary condition, and the effective air temperature will de­

pend on wind velocity and solar radiation.

To illustrate the influence of each significant variable on the allowable

heat input rates for steady state conditions with a limiting cable sheath

temperatures of 60° C and 90° C, a parameter study was done, with all compu­

tations referenced to a "standard" trench and cable system as follows:

Trench width 0.61 m (2.0 ft)


Trench depth 1.22 m (4.0 ft)
Depth of cable 0.91 m (3.0 ft)
Backfill resistivity, p 50 thermal ohms
Natural ground resistivity, Po 150 thermal ohms
Cable size 203 mm x 203 mm (8 in x 8 in) square
Ambient ground temperature, T 20° C
Air temperature, T^ 20° C
Wind velocity 8.05 km/hr (5 m.p.h.)
Solar radiation 0

For these conditions the allowable heat input rate to limit the cable

sheath temperature to 60° C is 105.8 watts per meter (32.25 watts per foot)

and to 90° C is 185.4 watts per meter (56.5 watts per foot).

5-28
The effect of solar radiation at the ground surface and the forced con­

vection owing to wind are accounted for, according to Barber (1957), in the

following way.

The effective air temperature is given by:

T (49)
E
where:

T effective air temperature, °C


E
T air temperature, °C
a
b absorptivity of the surface to solar radiation

- 0.65 for sand and gravel,

I solar radiation, cal/cm2/min.

Solar radiation at ground surface is approximately equal

to 1 cal/cm2/min.

h Convection coefficient, cal/cm2/min/°C

Convection coefficient is calculated from the relationship:

h = 1.3 + 0.62(V )0'75, BTU/ft2/hr/°F


W
where V = wind velocity, miles/hr
w

The relative importance of the several parameters influencing heat trans­

fer away from the buried cable are illustrated as follows.

1. Soil Properties

a. Backfill soil

The dominating influence of the thermal resistivity of the trench backfill

is illustrated by Fig. 50. The allowable heat input rate increases more than

proportionally with decrease in the backfill resistivity. This figure pro­

vides excellent illustration of the potential benefits to be achieved through

development of backfill materials with low thermal resistivity.

5-29
=20 C, Y"=8.05 km/hr. No radiation
Cobie sheath temp.

W o tts /m
60 C

40 60 60 lOO 120
Resistivity of Backfill Soil (Thermal ohms)

FIG. 50 EFFECT OF RESISTIVITY OF BACKFILL SOIL


b. Native soil

As would be anticipated, the lower the thermal resistivity of the native

soil, the greater the allowable heat input from the cable, as shown in Fig, 51,

In practice, however, the native soil is not a separately controllable variable,

2. Geometrical Factors

a. Cable Cross Section Size

Fig. 52 shows that the larger the cable size the greater the amount of

heat that can be dissipated.

b. Trench Depth

Fig. 53 shows that increasing the trench depth, for a given depth of cable

placement has a relatively minor effect on the allowable heat input rate.

c. Cable Depth

Fig. 54 shows that placing the cable at shallower depths increases the

allowable heat input rate.

d. Trench Width

Increasing the trench width from 0.61 m to 1,22 m (2 ft to 4 ft) gives an

increase in the allowable heat input rate of only about 20 to 25 percent for

trench depths in the range of 1.22 m to 1.83 m (4 to 6 feet).

e. Cover Thickness

Fig. 55 shows that the thicker the native soil cover placed over the trench

backfill the less the allowable heat input rate. This reduction amounts to

about 33 percent per meter (10 percent per foot) of cover for the conditions

shown.

3. Ambient Temperatures

a. Native Soil Temperature

Fig. 56 shows that ambient soil temperatures in the range of 0 to 30° C

have relatively little influence on the allowable heat input rate for a

5-31
Ta = 20°Ct Vy- 8.05km/hrt No radiation
Coble sheath temp.

5 80

203mm *203mm
^ 60

W atts/m
0 200 300
Resistivity of Native Soil (Thermal ohms)

FIG. 51 EFFECT OF RESISTIVITY OF NATIVE SOIL


mm
TA = 20 C, vw-8.05 km/hr. No radiation
”1 i/ojQ/ji r~ Cable sheath temp.
I O.S/5m A,‘ISO/l
1.22m „r/&s (S')
(4'i ^ |
0 .
T0‘£OC
90°C ^ 200
t
as
tf
/)
t

150
n
tIpt
uRt
a W
e(

t
t.
a

100

a
e

w
l
Alw
oal
beH

50

Square Cable Side Length (inches)


FIG. 52 EFFECT OF CABLE SIZE

5-33
m

Cable sheath temp.


90 °C -- ---

Tt *20*0, V*- 8.05 km/hr, No roOiatioti

0 915m p ’130/1

vanes ,3? T0--eo‘c

0.61m
• ■-i
li')

W a tts /m

60 C

4 5 6
Depth of Trench (feet)

FIG. 53 EFFECT OF TRENCH DEPTH


ON ALLOWABLE HEAT INPUT

5-34
m
LOO 1.25
~T" T"
Allowable Heat Input Rate (Watts/ft)

Watts/m
too

P‘50L
1.2Bm c varies
d
(4')

P Pq =150/1
50

T0 - 20°C
. tD3mm*£03 am *(2')

0 1 ___ I__________ l
2 3 4
Coble Depth (feet)

FIG. 54 EFFECT OF CABLE DEPTH


ON ALLOWABLE HEAT INPUT

5-35
m
600 0.25 0.50

Coble sheath temp.


Allowable Heat Input Rote ( Watts/ft)

- 150

- IOO

W a tt s /m
Ta =20°Cy Vw=8.05km/hr, No radiation

0.915 m
(3')
^ Po-tSOIL -50
v T0 " 20°C
0.6!m 203mm x203mm

Cover Thickness (feet)

FIG. 55 EFFECT OF COVER THICKNESS


ON ALLOWABLE HEAT INPUT

5-36
70

Coble sheath temp. =


- 200

8.05 km/hr, No radiation

0.915 m Pq =15011

W a tts /m
(S')
\ rn vanes -150

- 100

Ambient Soil Temperature (°C)

FIG. 56 EFFECT OF NATIVE SOIL TEMPERATURE


ON ALLOWABLE HEAT INPUT
steady-state cable temperatures of 60° and 90° C.

b. Air Temperature

The higher the air temperature above the ground surface the lower the al­

lowable heat input rate, Fig. 57. As high air temperatures and peak power

demands are likely to coincide it follows that system design may be controlled

by a high air temperature condition.

4. Wind Velocity

In the absence of solar radiation (e.g., at night) the magnitude of the

wind velocity is a relatively minor factor, as shown by Fig, 58.

5. Solar Radiation

Heat input to the ground can have a severe limiting effect on the ability

of the trench and backfill system to dissipate heat, as may be seen in Fig. 59

For incoming solar radiation in the amount of 1 cal/cm2/min, which is typical

for the ground surface on a sunny day, the allowable heat input rate is re­

duced by more than 30 percent of its value under conditions of no radiation.

In real installations alternating periods of night and day as well as cloud

cover should permit higher heat input rates for an "effective" solar radia-

tion less than 1 cal/cm /min.

6. Allowable Cable Surface Temperature

The allowable heat input rate increases in direct proportion to the maxi­

mum allowable cable sheath temperature as shown by Fig. 60.

7. Temperature Contours

Fig. 61 shows the steady-state temperature distribution around a cable

with a sheath temperature of 90° C and the standard geometry and backfill

conditions.

5-38
Tm varies, V„ -805km/hr, No radiation
Coble sheath temp.
Allowable H eat Input Rate (W a tts /ft)

200

s
t/m
150

a
Wt
100

20
Air Temperature (°C)

FIG. 57 EFFECT OF AIR TEMPERATURE ON ALLOWABLE HEAT INPUT


km/hr
Cable sheath temp.
<zn°fi _

Ta - 20 °C, varies. No radiation

/m
-40

s
0.61m^

tt
*(2')'

a
W
■s 30

20 —

Wind Velocity (miles/hr)

FIG. 58 EFFECT OF WIND VELOCITY ON ALLOWABLE HEAT INPUT


Ta =20*C ,Vw- 8.05 km/hr, / varies
0.915m
(S')

Cable sheath temp. — 150

Watts/m
60 C
§ 20

Solar Radiation (cal/cm2/min.)


FIG. 59 EFFECT OF SOLAR RADIATION
ON ALLOWABLE HEAT INPUT

5-41
Ta =20 C, Hg-8.05 km/hr, No radiation

0.9/5 m
(T)

W a tts /m
203 mm *203mm

60 -

10 60 80 100 120
Maximum Allowable Cable Surface Temp. (°C)

FIG. 60 EFFECT OF MAXIMUM ALLOWABLE CABLE SURFACE


TEMPERATURE ON ALLOWABLE HEAT INPUT
T =20°C, ^ - fl 05 km/hr, No radiation

6.1m
-43

(20')
§

s4s*C

20°C

6.! m
(20')

FIG. 61 TEMPERATURE CONTOURS FOR A CABLE AT 90°C


8. Two- and Three-Cable Systems

Fig. 62 shows the allowable heat input rates to limit cable sheath tem­

perature to 90° C for two- and three-cable systems in a single trench. It

may be seen that the allowable loading per cable is decreased for the multi­

cable arrangements.

9. Surface Cover

Cable systems are often buried under paved surfaces. Fig. 63 shows a

comparison between limiting heat input rates for trenches with native soil,

asphalt concrete, and Portland cement concrete in the top 152.4 mm (6 inches).

The differences are small, with the highest allowable heat input rate for the

Portland cement concrete because of its lower thermal resistivity.

Transient Conditions

Transient heat flow analyses have also been made. Fig. 64 shows the

time required for the cable surface temperature to reach 60° C and 90° C as a

function of heat input rate for the conditions shown. It may be seen that

several days would be required, even if there were a steady heat input at the

ground surface of 1 cal/cm2/min by radiation, which, of course, there couldn't

be because of night time conditions.

To examine further the increase of temperature as a function of time.

Fig. 65 was developed. The bottom portions of this figure show cable heat

input rate, effective air temperature, and solar radiation as functions of

time. In this case, the effective air temperature is considered to vary

sinusoidally and the temperature at any time t is given by:

2lTt
te - tm + Tv sin ir l50)

5-44
Ta 120°C, Vw = 8.05 km/hr, No radiation

T
_L \ \o9iSm P001150 sl
vanes
Tn=20°C

0.6!m 203mm*203mm

Ta = 20 C, Vw= 8.05 km/hr. No radiation

ph varies
1.525m

1.525m '203mm *
203mm

150

Ta = 20oC, Vw 18.05km/hr. No radiation

/>b varies

□ □Q
305mm* 305mm

20 40 60 80 100
Resistivity of Backfill Soil, Thermal Ohms

FIG. 62 ALLOWABLE HEAT INPUT RATES FOR MULTI-CABLE SYSTEMS

5-45
TA=2(fC, Vw*8.05 km/hr,No radiation
'24m mm^ ^//////////////M/i. ~ '"
(6304.8^nCrtls*?l*,‘* 9 ^
(I2"J j'™ Zf2Cfk

Cable Sheath

0.6lm

p=40Ji-
r Portland Cement Concrete
A llow able H eat In p u t R a te (W a tts /ft)

p=80Ji.
f/M/f/k-- Asphalt Concrete
A
-Native Soil
□ p-ISOn.

W a tts /m
Cable Sheath Temp.

Resistivity of Backfill Soil, pb (thermal ohms)

FIG. 63 ALLOWABLE HEAT INPUT RATE FOR DIFFERENT BACKFILL


MATERIALS IN THE TOP 6!! OF THE TRENCH

5-46
I

Wotts/m
FIG. 64 TIME REQUIRED TO REACH INDICATED TEMPERATURE
AS A FUNCTION OF HEAT INPUT RATE
rA *25%C, Vw-e.05 km/hr. Ivories
omsm fio1 'SOM.

-600

Watts/m
2 days 3 days 4 days
Time (hours)

FIG. 65 CABLE HEATING DURING A HYPOTHETICAL HEAT WAVE


where:

T = effective air temperature, °C

= mean effective air temperature, °C

= T + R
A
= average air temperature, °C
2 bX
R = — — , °C, b, I, and h are as defined previously

Ty = maximum variation in temperature from the mean. In the

day light hours;

T = 0.5 T + 3R, °C
V R
T = daily range in air temperature
R
Otherwise:

T 0.5 T , °C
V
t time, hours

The input parameters have been chosen to represent the conditions that

might accompany a summer heat wave. It was assumed for the analysis that the

trench backfill resistivity remained constant at 50 thermal ohms. Although

the plots in Fig. 65 are not real in the sense of pertaining to an actual

case, the results are useful in the sense that they demonstrate that by means

of the computational techniques now available it should be possible to anal­

yze the temperature vs. time behavior of buried cables for varying weather

and soil property conditions. Verification of these methods using actual

field data is desirable.

Heat Removal by Buried Thermal Conductors

The potential for heat removal by means of buried thermal conductors has

been analyzed. Aluminum sheeting has been considered as a buried radiation

5-49
material*. Table 7 compares three possible configurations relative to a trenc

in a natural ground having a thermal resistivity of 150 ohms. The cases of

both a natural soil backfill (p^ = 150) and a thermal sand backfill (p^ = 50)

are considered. The allowable heat input rates to limit the cable sheath tem­

perature to 60° C and to 90° C are shown for two cases of solar radiation

(1=0 and I = 0.5 cal/cm2/min). Times to reach these temperatures are also

indicated.

The following conditions are shown in Table 7.

1. No buried conductors.

2. Vertical strips underneath and vertically alongside the cable.

These strips cover half the vertical area along the line of the cable

3. Continuous conductor sheets underneath the cable and up the trench

sides.

4. Continuous conductor sheets along the trench sides and beneath and

above the cable.

It is clear from the results in Table 7 that substantial increases in

allowable heat input or reduction in cable temperature are possible if buried

thermal conductors are used. An arrangement such as shown for Case 2 could

be used to improve the thermal characteristics of existing cable systems.

The economics of systems such as shown in Table 7 have not been evaluated.

♦Because of corrosion, aluminum would not be suitable in many cases. Gal­


vanized steel might be a better choice. The principles established by the
examples should be valid in cither case.

5-50
Table 7 - Heat Removal by Buried Thermal Conductors

Q = Heat Input Rate


Ts = Cable Sheath Temperature
I = Solar Radiation Backfill Condition
1 2 3 4
= 50 n, 1=0:
Qall for Ts = 60oC (watts/m) 105.80 134.75 144.26 154.10
Qall" for Ts = 90°C (watts/m) 185.25 235.77 252.39 269.64

= 50 , I = 0.5 cal/cm2/min.:
Pb '
Qall. for Ts = 60OC (watts/m) 81.97 104.59 110.75 117.08
for T = 90°C (watts/m) 158.39 202.03 214.00 226.23
Time'to Ts = 60°C for Q = 295.1 watts/m 2.26 2.98 3.25 4.18
(90 watts/ft) (days)
Time to Ts = 90°C for Q = 295.1 watts/m 9.00 13.87 16.90 23.14
in
i p = 150 Q, 1=0: (90 watts/ft) (days)
in
i—1
Qall> for T^ = 60°C (watts/m) 60.66 93.11 103.31 11^.21
Qall. for Ts = 90°C (watts/m) 106.16 162.92 180.82 208.62

p = 150ft, I = 0.5 cal/cm2/min.:


Qall for Tg = 60°C (watts/m) 48.52 75.74 82.98 94.16
Qan for Tg = 90°C (watts/m) 93.84 146.49 160.46 182.07
Time to T = 60°C for Q = 295.1 watts/m 1.20 1.95 2.47 3.81
S (90 watts/ft) (days)
Time to Ts = 90°C for Q = 295.1 watts/m 3.51 6.65 9.09 14.10
KEY SKETCHES: ^90 watts/ft) (days)
Ta =20C, V*-8.05 km/hr Ta=20% Vw=a05 km/hr Ta-20C, Vw= 8.05 km/hr
Ta=20°Ci \4f=B05ttm/f>r, I varies
I varies Ivories I varies
PjISOJi. J
P.
vanes r
0.9/m T I
If PfflSOJL
asm T=2(fc
TpJI fif*150 I Poxl5°
\0.9/m t0*BO°C
T0*20*C L22m
Tl* t° *oc
L
\ 203mm*
iiLJ8 mm Aluminum 3.18mm Aluminum 338 trm Aluminum
203mm o.eim\ "p varies

CASE 1 CASE 2 CASE 4

NOTE: Continuous aluminum sheets assumed in longitudinal direction for Cases 3 and 4.
VI. MOISTURE MIGRATION UNDER THERMAL GRADIENTS

Introduction

The phenomenon of water movement in soils due to thermal gradient has

been studied in relation to several problems, including the availability of

soil water for plant growth as a function of soil temperature changes due to

diurnal and seasonal temperature cycles, and the moisture accumulation in high­

way pavement layers which could adversely affect the strength of the pavement.

In underground power cable systems temperature gradients are developed around

the cable. These thermal gradients induce soil water movement away from the

cable, which, in turn, can lower the thermal conductivity of the cable environ­

ment causing an undesirable build up of heat around the cable.

Thermally induced soil water movement in unsaturated soils is a complex

phenomenon. The various components of soil water movement and the physical

laws governing them are presented in this section. The concepts, relation­

ships, and data presented will be used as a basis for subsequent analyses and

predictions of moisture migration in buried cable systems.

Statement of the Problem

The problem of moisture movement under thermal gradients in unsaturated

soils is less understood than its movement in saturated soil. This is due to

the simpler transport mechanism involved with saturated flow as opposed to

unsaturated flow. In saturated soils the moisture movements occur in the

liquid phase only. In unsaturated soils the flow takes place in two phases,

vapor and liquid.

A comprehensive review of the literature on water migration under thermal

gradients was made. This review showd that several theories now exist. The

theories of Taylor and Cary (1964) and Philip and DeVries (1957) have received

6-1
the most attention because they are the most rigorous. Laboratory experiments^

Cassel et. al. (1969), showed close correlation with the Philip and DeVries

theory while Taylor and Cary's approach underestimated the flow by 10 to 40

fold. Therefore, the theory of Philip and DeVries appears to provide the most

comprehensive basis for development of a model for prediction of transient flow

of water in unsaturated soils under thermal gradients. This theory has ten­

tatively been selected for analysis of the buried cable problem. Both theories

are summarized in the following paragraphs.

Taylor and Cary Theory

Taylor and Cary (1964) used the general theory of irreversible thermo­

dynamics to account for the simultaneous flow of water and heat in porous

media under the influence of thermal gradients and other secondary energy

gradients. The porous medium is considered as a real system with coupling

flow phenomena leading to a net increase in the entropy of the system.

Taylor and Cary developed a linear flow equation for each component of the

soil system which has the following general form:

r • • • • r n) (51)
k=l

where:

is the flux of the i-th phenomenon (e.g. vapor transport or

energy transfer).

is the driving force k for any irreversible process, such as heat

and concentration gradient.

is the phenomenological coefficient describing flow in the i-th

phenomenon due to the action of driving force X^, such as the dif­

fusion coefficient and hydraulic conductivity.

n is the number of driving forces.

6-2
According to Onsager's reciprocal relationship:

Lik = \i

Taylor and Cary's equation which describes steady thermal moisture movement

is as follows:

d(lnT)
+ B (52)
dx

where:

D is the water diffusivity

^ and are the gradients of moisture and the logarithm of

temperature, respectively

B is the transfer coefficient

_ _ d9 _ Lwq
d(lnT) D

Lwq is the phenomenological coefficient for thermally induced water

transfer with subscripts w and q referring to water and heat.

B was assumed to be constant to satisfy the linearity requirement of

the flow equation: Taylor and Cary studied the data of Taylor and Cavazza

(1954) and found that B is constant only in moist soils. Gee (1966), Cassel

et. al. (1969) observed that B is dependent on moisture content range and

temperature.

Philip and DeVries Theory

In an analysis of moisture movement in unsaturated soils under a thermal

gradient the following must be considered:

- The vapor phase trasnport

- The liquid phase transport

6-3
- The interrelationship of the vapor and liquid phases.

Vapor Phase Transport. The equation of vapor diffusion, modified so as to

apply to porous media. Penman (1940), may be written as:

q = -D vaaVp (53)
vap o

where:
O
the vapor flux density, gm/cm /sec
2
D is the molecular diffusivity of water vapor in air, cm /sec
o
a is a tortuosity factor allowing for extra path length

3 3
a is the volumetric air content of the medium, cm of air/cm

p is the density of water vapor, gm/cm

V is the "mass flow factor" introduced to allow for the mass flow

of vapor arising from the difference in boundary conditions

governing the air and vapor components of the diffusion system.

It is equal to P/(P - p) where P is the total gas pressure and p

is the partial vapor pressure in the pore space.

Equation (53) has been called the simple theory of vapor transfer. It

can be extended to separate the isothermal and thermal components of vapor

transfer and to take into account the effect of relative humidity on the

trasnfer. Introducing the thermodynamic relationship (Edlefsen and Anderson,

1943):

p = p h = p exp(il;g/RT) (54)
o o
where:

pQ is the density of saturated water vapor, gm/cm3

h is the relative humidity

g is the acceleration due to gravity, cm/sec

R is the gas constant of water vapor, erg/gm/°C

ijj is the suction head, cm, in thermodynamic equilibrium with the

6-4
water in the medium (iJj is negative in unsaturated soils)

T is the absolute temperature, °k

gives
Vp = hVp + p Vh (55)
o o

As p^ is a function of T only and - 0 in the full range of h, there­

fore, h is a function of 0 (volumetric water content) only.

Thus equation (55) becomes:

dp.
VPK = h-^r VT + p ve (56)
dT Ko d9

substitution for in (56) from (54) gives


dU

Vp h VT + p I! V0 (57)
dT o RT 30
substitution of (57) into (53) gives

/ dpo . P°g 3^ '


q = -D vaa ( h —— VT (58)
^vap o y dT + -iF W V0
or

-DTv VT - D0V V0 (59)


Srap^W
where:

is the density of water, gm/cm

is the thermal vapor diffusivity, cm /sec/°C


D dp
= — vaah -r2- (60)
p„ «
2
is the isothermal vapor diffusivity, cm /sec
D p g ^,
O O^ dip (61)
= — vaa —-
p rt 30
w

In this way the vapor flux has been separated into two components, that

due to temperature gradient and that due to moisture gradient.

Liquid Phase Transport. Darcy's law for liquid transfer in unsaturated media

may be written as:


6-5
-K0 V * (62)
qliq/PW
where:

q . is the liquid flux density, gm/cnr/sec

is the unsaturated hydraulic conductivity, cm/sec


0
<J> is the total potential, cm

The total potential <J> is composed of pressure and gravitational components

<() = ^ + Z (63)

where Z is the vertical ordinate, positive upwards.

In the moisture content range where liquid transfer occurs (K0 > 0) , if;

is determined by capillarity, hence:

(64)
9t 0 dT
where:

0 is the surface tension of water, dynes/cm

Substituting equations (63) and (64) into (62) yields:

qliq/PW -Ke lllVT - k9H,s - V (65)

where i is the unit vector in the positive Z direction. Rewriting:

q, . /prT = -D VT - Dq V0 - KQi (66)


^liq W T^ 0^ 0

where:

D_ is the thermal liquid diffusivity, cmz/sec/°C

= K, t 52. (67)
0 0 dT

is the isothermal liquid diffusivity, cm2/sec

(68)
*0 30
Thus, the liquid flux has been separated into three components, that due

to temperature gradient, that due to moisture gradient, and that due to gravity

6-6
Interaction of the Vapor and Liquid Phases. Many workers (Gurr et. al. , 1952

Taylor et. al., 1954, Rollins et. al., 1954) found that observed water vapor

transport under a temperature gradient greatly exceeds that predicted by the

simple vapor transport theory presented above.

Philip and DeVries modified the simple theory to account for the inter­

action between vapor, liquid and solid phases, and the difference between

average temperature gradient in the air-filled pores and in the soil as a

whole. With these factors taken into account, they developed an expression

which predicts order of magnitude and general behavior in satisfactory agree­

ment with the experimental facts. The modified expression for D is as


TV
follows:

D dP
D = -—V (a + f(a) • 0)h£ , cm2/°C/sec (69)
TV PW dT

where:

f(a) = a/a for 0<a<a


K K

= 1 for a>a
K

where a is the value of a (volumetric air content) at which


K.
liquid continuity in pores no longer exists, or a is the
K
value of a at 9 = 0 where 0 is the moisture content at which
K K
K (hydraulic conductivity) falls to some small arbitrary frac­

tion of its saturated value.

£ is the ratio of the average temperature gradient in the air-

filled pores to the overall temperature gradient.

Comprehensive Theory. The total water movement in an unsaturated soil due

to a temperature gradient and its resulting water content gradient is equal

to the sum of the flows occurring in the two phases, vapor and liquid. Thus:

6-7
(70)
q/pw ’ qvap/Pw + qUq/PW

2
where q is the total water flow, gm/cm /sec.

Substituting the expressions for qvap/Pwf


q equation (59) and q^^/Pw,

equation (66), in equation (70) yields:

q/p. (D + D ) VT - (D + D ) V0 - K i
rv l °V Z

—D VT - Dq V0 - K„i (71)
T 0 0

where:

dt D + D Thermal Water Diffusivity (72)


T£ TV

Isothermal Water Diffusivity (73)


D0 Dfl + Dfl
0£ 9V

Equation (71) is the governing law for moisture movement under thermal

gradient in unsaturated soils as proposed by Philip and DeVries.

Differentiation of (71) and application of the continuity requirement

gives the general differential equation:

9k,
(74)
H - V(dtVt) + v<D07e> + ~W

The heat conduction equation for the soil is:

c = V(AVt)

If -V(?VT)

3t (75)
V(xVt)
3t

where:

C is the volumetric heat capacity of the soil, cal/cm /°C

X is the thermal conductivity, cal/cm/sec/°C


o
X is the thermal diffusivity, cm /sec

6-8
Equations (74) and (75) are the simultaneous equations describing coupled

heat and moisture transfer in partially saturated soils.

Evaluation of the Soil Parameters

From equations (74) and (75) it is clear that the variation of four soil

parameters with the water content (or with the degree of saturation) is needed

in order to predict the moisture transfer. These parameters are:

Kg Hydraulic conductivity

Dg Isothermal water diffusivity

Dt Thermal water diffusivity

X Thermal diffusivity

These four parameters have been evaluated for three different soils; the

Fire Valley thermal sand, Round-Robin sand, and Monterey #0 sand. The tem­

perature range considered for evaluation of these parameters was 10-60° C.

The details of the methods of evaluation and the results obtained are given

below.

Hydraulic Conductivity, Kg. Kg can be obtained by either of the following two

methods (the second method was used by us in evaluating Kg and it is presented

in more detail here):

1) A vertical soil column experiment in which the water content and the

potential gradient are measured as a function of time, and Kg is obtained

from the following continuity equation:

30 _ 3q
3t - 3z;

= J_ (K (76)
3s K0 3z

where:

0 is the volumetric water content, cm3/cm3

6-9
t is the time, sec.

q is flow rate, cm3/sec

is potential gradient.

2. Marshall (1958) derived an equation for the relation between the hydraulic

conductivity and the size distribution of the pores in isotropic material,

K = e2 n"2[r2 + 3r2 + 5r2 + . . + C2n - l)r2]/8 (77)


12 3 n

where:
2
K = specific hydraulic conductivity, cm

E = porosity, cm3/cm3

n = total number of pore classes

r^ = mean radius of the pores in the pore class i, cm

Equation (77) makes it possible to calculate the hydraulic conductivity

of porous material from the size distribution of its pores. Pore size

measurements can be made by measuring the water withdrawn when the suc­

tion on the water is progressively increased, 'fhe radius of the largest

pores full of water when the suction of ip cm is applied is given by:

2a
(78)
pgip

where:

a = surface tension of water, dynes/cm

p = density of water, gm/cm


2
g = acceleration due to gravity, cm/sec

In applying equation (77), it is often more convenient to use the suction

in place of the pore radius. Therefore, equation (77) becomes:

2
K = -2-y-j e2 n_2[^2 + 3ip~2 + 5ip~2 + . . . + 2(n - Dip"2] (79)
2p g

6-10
The units of the hydraulic conductivity can be transferred to cm/sec by

multiplying the specific hydraulic conductivity by Pg/H where p and g

are as defined above and p(gm/cm/sec) is the viscosity of water. Thus,

we obtain:

2
K(cm/sec) = -- £2 n 2 [tit2 + 3il; 2 + 2 + . . . + 2(n - 1)^ 2] (80)
2 pgn 12 3 n

In the derivation of this equation by Marshall, £ was used as the water-

filled porosity at each water content 0^ for which the hydraulic conduc­

tivity (K) is calculated, and n was the number of pore classes in the

water content interval from zero to 0.. Thus both £ and n decrease as
i
the water content decreases. Green and Corey (1971) proposed another

approach which is consistent in concept and leads to the same results.

They assumed £ to be the water-saturated porosity, 0s, and n to be the

total number of pore classes at £ = 0^. In this approach £ and n remain

constant when the hydraulic conductivity (K) is calculated for different

values of water content (0). This approach was used by us in calculating

the hydraulic conductivity of our soils.

Rewriting equation (80) in the following form and introducing a matching

factor which is usually the ratio of the measured to calculated hydraulic

conductivity at full saturation, we obtain:

2
2 t(2j + 1 - 2i) ^2] (81)
K(9)i = K- ^ ^
sc n
j=l i-l,2,...,m

A summary of the definitions of the terms used in equation (81) and their

units is as follows:

6-11
K(0) . calculated hydraulic conductivity for a specified water
i
content, cm/sec

e volumetric water content, cm /cm

the last water content class on the wet end, e.g., i = 1

denotes the pore class corresponding to the saturated water

content, and i = m denotes the pore class corresponding to

the lowest water content for which hydraulic conductivity is

calculated.

matching factor (measured saturated conductivity/calculated

saturated conductivity

a surface tension of water, dynes/cm

P density of water, gm/cm3


2
g acceleration due to gravity, cm/sec

n viscosity of water, gm/cm/sec

e water-saturated porosity, cm /cm

n total number of pore classes between 0=0 and 0 where 0


*D
is the water content at saturation.

m total number of pore classes between 0=0 and 0 where 0


■L Ij
is the lowest water content on the experimental moisture

characteristic curve. Thus:

n m
0 0
S L

= suction head (negative pressure head) for a given class of water

filled pores, cm of water.

m is constant at all water contents, and the value of m establishes the


-2
number of pore classes for which lb. terms are included in the calcula-

tion at saturation, m should be chosen so that the water content-suction^


4
6-12
head relationship is represented accurately. A value of m = 12 was chosen

to calculate the results given in this report.

From the comparison between experimental and calculated values, it has

been concluded (Green et. al. (1971), Kunze et. al. (1968), Elzeftawy et. al.

(1977), Nielsen et. al. (I960)) that this method, after introducing the matching

factor, successfully predicts the hydraulic conductivity for a wide range of

soils.

The suction head-water content relationship was obtained for each of our

soils by using the commercially available "Tempe" pressure cells, which are

suitable for the low range of pressure, 0-1 atm., needed here. After the soil

is placed in the cell, it is saturated with water. The soil is allowed to

drain following sequential subjection to increasing air pressure. Volumetric

water content is determined from the value of the withdrawn water after each

static equilibrium pressure and the water content of the sample at the end of

the test. The suction head-water content relationships obtained for Thermal

sand. Round-robin sand, and Monterey #0 sand are shown in Fig. 66 through

68, respectively. The suction heat-water content is a hysteretic function;

i.e., 0 depends upon whether the soil is drying or wetting. The data in Figs.

66-68 are for drying.

The saturated hydraulic conductivity for various void ratios (dry densi­

ties) was measured for each soil in the laboratory using the falling head per­

meability test equipment. Equation (81) was used to calculate the hydraulic

conductivity for each soil using a short computer program written for this

purpose. The hydraulic conductivity versus degree of saturation plots for

Fire Valley thermal sand. Round-robin sand, and Monterey #0 sand are shown

in Figs. 69 through 71, respectively.

6-13
Dry Unit Weight =19.5 kN/rrP(!24pcf)
o f water)
#(cm
Suction Head ,

Volumetric Water Content, 0 (cm3/cm3)

FIG. 66 VOLUMETRIC WATER CONTENT-SUCTION HEAD


RELATIONSHIP FOR FIRE VALLEY THERMAL SAND

6-14
1000
Dry Unit Weight = 20.4 kN/m303Opcf)
o f w ater)
(cm
\fr
Suction Head,

Volumetric Water Content, 0 (cm3/cm3)

FIG. 67 VOLUMETRIC WATER CONTENT-SUCTION HEAD


RELATIONSHIP FOR ROUND-ROBIN SAND

6-15
Dry Unit Weight =16.0 kN/m3 002pcf)

O.i 0.2 0.3


Volumetric Water Content, 0 (cmS/cm5)

FIG. 68 VOLUMETRIC WATER CONTENT-SUCTION HEAD


RELATIONSHIP FOR MONTEREY NO. 0 SAND

6-16
Dry Unit Weigh t=!9.5 kN/m3 (124pcf)
K*/KCr =0.047

40 60 SO 100
Degree of Saturation, S (%)

FIG. 69 VARIATION OF HYDRAULIC CONDUCTIVITY OF FIRE


VALLEY THERMAL SAND WITH DEGREE OF SATURATION

6-17
Dry Unit Weigh*=20.4 kN/m3030pcf)

Degree of Saturation, S (%)

FIG. 70 VARIATION OF HYDRAULIC CONDUCTIVITY OF ROUND-ROB


SAND WITH DEGREE OF SATURATION

6-13
H yd rau lic Conductivity, K# (c m /s e c .) Dry Unit Weight =16.0kN/m3002pcf)

Degree of Saturation, S (%)

FIG. 71 VARIATION OF HYDRAULIC CONDUCTIVITY OF MONTEREY


NO. 0 SAND WITH DEGREE OF SATURATION

6-19
Isothermal Water Diffusivity, D, From equation (73)
—e*

\ + \

where:

Isothermal liquid diffusivity, cnr/sec.

Dq = Isothermal vapor diffusivity, cm/sec.


9V

(a) D
e,

From equation (68)

0, 0 80

Since the soil water content-suction head relationship is hysteretic,

it follows that DQ is also hysteretic function. Then, in calculating DQ from


dSL
Kg data, it is necessary to consider whether the soil is wetting or drying.

The values of DQ , for the drying process of the soil, versus degree of sat-

uration for Thermal sand. Round-robin sand, and Monterey #0 sand are shown in

Figs. 72 through 74, respectively.

(b) D,
0.

From equation (61)

D P g n. |
„ o o
\ ' ^vaa^
D : Krischer and Rohnatter (1940) found that D , the diffusion co-
o o
efficient for water vapor in air, could be represented by the
_4 ip2.3
expression: 4.42 X 10 —--- in the temperature range 20-70° C,

where T is the absolute temperature and P is the total gass ^

pressure, mm Hg.

6-20
Dry Unit Weight =19.5 kN/m3(124pcf)

$ 10 —

Degree of Saturation, S (%)

FIG. 72 VARIATION OF ISOTHERMAL LIQUID DIFFUSIVITY OF FIRE


VALLEY THERMAL SAND WITH DEGREE OF SATURATION

6-21
Dry Unit Weight= 20.4 kN/m3 (130pcf)
(cm/sec.)
Isothermal Liquid Diffusivity, D a

> 40 60 80
Degree of Saturation, S (%)

FIG. 73 VARIATION OF ISOTHERMAL LIQUID DIFFUSIVITY OF


ROUND-ROBIN SAND WITH DEGREE OF SATURATION

6-22
Dry Unit Weight = 16.0 kN/m3 (102pcf)
Isothermal Liquid Diffusivity, Da. (cm 2/sec.)

Degree of Saturation, S (%)

FIG. 74 VARIATION OF ISOTHERMAL LIQUID DIFFUSIVITY OF


MONTEREY NO. 0 SAND WITH DEGREE OF SATURATION

6-23
V: V = P/(P - p) = mass flow factor

P is the total gas pressure (atmospheric pressure in the case

of soils)

p is the saturation vapor pressure.

a: Penman (1940) gave the tortuosity factor, a, a value of 0.66.


0
R: Gas constant = 4.615 X 10 erg/gm/°C

Isothermal vapor diffusivity versus degree of saturation relationships

for Thermal sand, Round-robin sand, and Monterey #0 sand are shown in Figs. 75

through 77, respectively.

Thermal Water Diffusivity, PT. From equation (72)

D + D
T£ TV

where:

Thermal liquid diffusivity, cnr/sec/°C

Thermal vapor diffusivity, cnr/sec/°C

(a) D

From equation (67)

, i da
^0 a dT

The surface tension temperature coefficient can be taken as a


-3
constant value of -2.285 X 10 /°C in the temperature range 10-60° C. Thus:

-3 9
D = -2.285 X 10 K„i{j, cm /sec/°C
T£ 9

6-24
Dry Unit Weight = 19.5 kN/m3(124pcf)
Is o th e rm o l V o p o r D iffu s iv ity , D^v (c m /s e c .)

Degree of Soturotion, S (%

FIG. 75 VARIATION OF ISOTHERMAL VAPOR DIFFUSIVITY OF FIRE


VALLEY THERMAL SAND WITH DEGREE OF SATURATION

6-25
(c m /s e c .)
Is o th e r m o l V o p o r D iffu s iv ity , D a Dry Unit Weight =20.4 kN/m3(I30pcf)

Degree of Soturotion, S (%)

FIG. 76 VARIATION OF ISOTHERMAL VAPOR DIFFUSIVITY OF


ROUND-ROBIN SAND WITH DEGREE OF SATURATION

6-26
Dry Unit Weight = 160 KN/m3 (102 pcf)
Is o th e r m a l V o p o r D iffu s iv ity , Dg^ (c m y s e c .)
2

40 60 80
Degree of Saturation, S (%)

FIG. 77 VARIATION OF ISOTHERMAL VAPOR DIFFUSIVITY OF


MONTEREY NO. 0 SAND WITH DEGREE OF SATURATION

6-27
(b)

From equation (69)

D dp
D = — v (a + f(a) • 6)he —£
T p dT
V MW

h: the relative humidity, h, was evaluated from the thermodynamic

relationship presented in equation (54):

h = exp(i|jg/RT)

£: e = (Vt) /Vt
3l

(Vt) is the average temperature gradient in the air-filled pores

(Vt) is the overall temperature gradient

e was evaluated in the manner illustrated by DeVries (1963).

8p 8p
-g^-: the change of vapor density at saturation with temperature,

was obtained from physical chemical tables for the temperature

range 10-60° C.

The thermal water diffusivity versus degree of saturation relationships

for Thermal sand. Round-robin sand, and Monterey #0 sand are shown in Figs. 78

through 80, respectively.

Thermal Diffusivity, X. ’The thermal diffusivity is given by:

X = | (82)

The thermal conductivity, , was evaluated, for the three soils, in the

laboratory using the thermal needle method.

The volumetric heat capacity, C, was obtained from the expression:

Y, (C + •) , cal/cmV°C
100

6-28
Dry Unit Weight = 19.5 kN/m3 (124 pcf)
T h erm al W ater D iffu s iv ity , D T (cm 2/°C /s e c .)

vopor

Degree of Saturation, S (%)

FIG. 78 VARIATION OF THERMAL WATER DIFFUSIVITY OF FIRE


VALLEY THERMAL SAND WITH DEGREE OF SATURATION

6-29
Dry Unit Weight = 20.4 kN/m3 (130pcf)
T h e rm a l W ater D iffu s iv ity , D j (c rrfy ^ c /s e c .)

vapor

Degree of Saturation, S (%)

FIG. 79 VARIATION OF THERMAL WATER DIFFUSIVITY OF


ROUND-ROBIN SAND WITH DEGREE OF SATURATION

6-30
( c m /^ /s e c .)
T h erm o l W oter D iffu s iv ity , D T

Dry Unit Weight =16.0 W^m3 (I02pcf)

vapor

Degree of Saturation, S (%)

FIG. 80 VARIATION OF THERMAL WATER DIFFUSIVITY OF


MONTEREY NO. 0 SAND WITH DEGREE OF SATURATION

6-31
where:

C = specific heat of dry soil

= unit dry weight, gm/cm3

W = water content, percent dry weight.

The thermal diffusivity versus degree of saturation relationships for

Thermal sand, Round-robin sand, and Monterey #0 sand are shown in Figs. 81

through 83, respectively.

Planned Application

The relationships and properties presented in this section make possible

the prediction of moisture movements under combined hydraulic and thermal

gradients. It is intended that they be applied for analysis of typical cable,

trench, and backfill systems. The complexity of the relationships and their

interdependence requires that computer solutions be used. Programs are avail­

able for analysis of flows in one dimension. The development of a program for

analysis of flows in two dimensions is being considered.

Predictions of temperature distributions and moisture movement will be

made in connection with a field test program to be conducted during the next

two years of research on this project. Measurements in the field will then

be compared with predicted values to ascertain the accuracy of the theory.


(cm^sec.)
x I0 3
T h e rm a l D iffu s iv ity , X

Dry Unit Weight = 19.5 kN/m3 (124 pcf)

Degree of Saturation, S (%)

FIG. 81 VARIATION OF THERMAL DIFFUSIVITY OF FIRE


VALLEY THERMAL SAND WITH DEGREE OF SATURATION

6-33
(cm2/s e c .)
X x |0 3
Therm al D iffu s iv ity ,

Dry Unit Weight = 212 kN/m3(l35pcf)

Degree of Saturation, S (%)

FIG. 82 VARIATION OF THERMAL DIFFUSIVITY OF ROUND-ROBIN


SAND WITH DEGREE OF SATURATION

6-34
Dry Unit Weight =16.0 kN/m3 (102 pcf)

Degree of Saturation, S (%)

FIG. 83 VARIATION OF THERMAL DIFFUSIVITY OF MONTEREY


NO. 0 SAND WITH DEGREE OF SATURATION

6-35
VII. SUMMARY AND CONCLUSIONS

The objective of the research described in this report is to develop im­

proved methods for placing backfill around underground power cable systems and

special treatments to reduce the thermal resistivity and increase the thermal

stability of the backfill materials. The scope of work accomplished in the

two years since initiation of research on this project has included several

studies aimed at meeting this objective. The principal findings and conclu­

sions are summarized below.

Measurement of Thermal Resistivity

Of the available methods for evaluation of the thermal resistivity of

soils of the type used as backfill materials around buried power cables, the

thermal needle method and the Shannon and Wells method have been used most

extensively and are considered most suitable because of their relative sim­

plicity and the short measurement time required. The thermal needle method

is the simpler of the two, and it offers the added advantage that knowledge

of the specific heat is not required for calculation of thermal resistivity

from the test data.

A special thermal needle design (UC thermal needle) has been developed,

and the influence of test conditions have been studied in detail. The test

procedures finally developed are simple, rapid, and give reproducible results.

The UC thermal needle makes possible the determination of thermal resistivity

in both the radial and axial direction for a given sample.

Effect of Compaction Method

Four different methods of sample compaction (tamping, pluviation, rod-

ding, vibration) were investigated to determine whether or not the different

7-1
grain arrangements (fabric) produced resulted in different values of thermal

resistivity. It was found that the method of sample preparation had no signi­

ficant effect on samples as compacted, provided comparisons were made at the

same sample density. On the other hand, the thermal resistivity of samples

contacted wet and then dried are substantially less than for samples compacted

dry.

Backfill Treatments

Additives which can either (1) prevent water migration in the soil due

to temperature gradients, or (2) substitute for water in bridging the gaps

between particles have potential application for buried cable backfill treat­

ment. A large number of different additive materials were tested to determine

their effectiveness. Criteria that must be considered in the selection of

materials to meet the above objective include cost, availability, toxicity,

stability and durability, ease of handling, required treatment levels, and

effectiveness with different soil types. Tests were done on both Monterey

No. 0 sand, a uniform sand of very high thermal resistivity (62 °C-cm/watt at

10 percent water content, 340 °C-cm/watt when dry), and Fire Valley thermal

sand, a well-graded sand and gravel mixture with low thermal resistivity

(about 50 °C-cm/watt when at 6 percent water content, 100 °C-cm/watt when

dry).

Of the inorganic materials tested portland cement and C-7, a cement-based

additive, were the most effective. They are relatively low cost; however,

they harden the backfill, thus making subsequent access to the buried cable

difficult.

Three water-absorbing and water-holding polymers were studied and found

to delay the rate of drying of treated backfill material. Thus the thermal

7-2
resistivity increase at a slower rate than for untreated material when dried

at elevated (60 °C) temperatures. Samples treated with these additives were

difficult to compact to high densities and to separate interparticle contact

points. Their use would likely be detrimental in materials that are subject

to complete drying.

Waxes and asphalt are the most promising of the additives studied thus

far in terms of effectiveness in reducing thermal resistivity, ease of incor­

poration and handling, and economics. Unrefined slack wax and emulsified waxes

are available at low cost and appear about equally effective. Preliminary

indications are that waxes can maintain the thermal resistivity of dry,

treated Fire Valley thermal sand at values of 50 to 60 °C-cm/watt indefinitely;

i.e., the waxes appear to be effective, permanent substitutes for water.

Durability tests are continuing to strengthen these conclusions.

Temperature Distributions Around Buried Cables

Methods for analysis of temperature distributions around buried power

cables were reviewed and evaluated. None of the previously published proce­

dures accurately modelled the true geometric conditions or properties. All

give steady-state temperature distributions, so that times for attainment of

equilibrium or the influence of cyclic thermal loading cannot be evaluated.

Consequently, the finite element computer program HEAT was used to study

both transient and steady-state temperature distributions both within the

cable trench backfill and the surrounding soil. A parameter study was made

to determine the relative importance of the several variables that influence

the rates of heat flow and temperature distribution. Allowable heat input

rates to limit cable sheath temperature to 60 °C and 90 °c were determined.

The thermal resistivity of the trench backfill has a dominating influence.

7-3
Allowable heat input increases significantly with increase in cable size.

Increasing the trench size beyond sizes in common use; e.g. for one cable

system, 0.61 m (2 ft) wide by 1.22 m (4 ft) deep, has a relatively minor in­

fluence. In the absence of solar radiation (e.g., at night) wind velocity is

relatively unimportant. On the other hand incoming solar radiation on a calm

day can have a severe limiting effect on allowable heat input.

The computational procedures allow for prediction of temperature as a

function of time for varying weather and soil property conditions. The re­

liability of these predictions will be evaluated by means of field tests during

the next phase of this project. Additional analyses show that substantial in­

creases in allowable heat input or reduction in cable temperature are possible

if buried thermal conductors are used.

Moisture Migration Under Thermal Gradients

Thermally induced water movement in unsaturated soils is a complex phe­

nomenon involving both vapor and liquid phase flow. The various components

of soil water movement and the laws governing them have been reviewed, and

the theory of Philip and deVries seems most suitable for prediction of tran­

sient water flow. To apply this theory the variation of four soil properties

(hydraulic conductivity, isothermal water diffusivity, thermal water diffusi­

vity, and thermal diffusivity) are needed as a function of water content.

These properties have been evaluated for three sands characteristic of those

to be encountered in buried cable installations. Analysis of moisture move­

ments in typical trench, cable, and backfill systems will be made using these

properties and the Philip and deVries theory. A computer program for analysis

of flows in one dimension is available; extension to two dimensions will be

done if possible.

7-4
Future Work

Research for the next two-year period is to include further laboratory

tests to evaluate new additives for reduction of thermal resistivity, further

evaluation of the durability and properties of wax-treated backfill materials,

evaluation of the effectiveness of wax with different representative soil

types, and a field test program. The field test program will simulate actual

buried cable installations, and it will provide a basis for comparison between

predicted and actual moisture and heat flows, as well as field evaluation of

the most promising backfill treatments.

7-5
REFERENCES

Barber, E.S., "Calculation of Maximum Pavement Temperature from Weather Reports,"


Highway Research Board, Bulletin 168, Washington, 1957.

Bauer, C. A., and Nease, R. J., "A Study of the super position of Heat Fields
and the Kennelly Formula as applied to Underground Cable Systems," AIEE
Trans, V, Part III, p. 1330-1337, February, 1958.

Carlslaw, H. S. and Jaeger, J. C., Conduction of Heat in Solids, 2nd Edition,


Oxford University Press, New York, 1959.

Cassel, D. K., Nielsen, D. R., and Biggar, J. W., "Soil water movement in
response to imposed temperature gradients," Soil Science Society of
America, Proc. 33: 493-500, 1969.

De Vries, D. A., "Thermal Properties of Soil". In W. R. Van Wijk (ed.);


"Physics of plant environment". J. Wiley & Sons, Inc., New York, 1963.

Edlefsen, N. E., and Anderson, A. B., "The thermodynamics of soil moisture,"


Hilgradia, 16, 31-299, 1943.

Elzeftawy, A. and Dempsey, B. J., "Prediction Model for Hydraulic Conductivity


and Moisture Diffusivity of Highway Soils," Paper presented at the 56th
Annual Meeting of the Transportation Research Board, January 1977.

Gee, G. W., "Water movement in soils as influenced by temperature gradient",


Ph.D. Thesis, Washington State University, 1966.

Green, R. E. and Corey, J. C., "Calculation of Hydraulic Conductivity: A


Further Evaluation of Some Predictive Methods," Soil Science Society of
America, Proc. 35: 3-8, 1971.

Gurr, C. G., Marshall, T. J., and Hutton, J. T., "Movement of water in soil
due to a temperature gradient," Soil Science, 74, 335-345, 1952.

Krischer, 0. and Rahnatter, H., "Warmeleitung und Dampfdiffusion in Feuchten


Gutern," Verein Deut. Ing-Forschungsheft 402, 1940.

Kunze, R. J., Vehera, G. and Graham, K., "Factors Important in the Calculation
of Hydraulic Conductivity," Soil Science Society of America, Proc. 32:
760-765, 1968.

Marshall, T. J., "A Relation Between Permeability and Size Distribution of


Pores," J. Soil Science, Vol. 9, 1958.

Mitchell, J. K., Fundamentals of Soil Behavior, John Wiley & Sons, Inc.,
New York, 422 p., 1976.

Neilsen, D. R., Kirkham, D. and Perrier, E. R., "Soil Capillary Conductivity:


Comparison of Measured and Calculated Values," Soil Science Society of
America, Proc. 24: 157-160, 1960.

R-l
Penman, H. L., "Gas and Vapor Movement in Soil," I, Journal Agr. Science, 30,
437-462, 1940.

Philip, J. R. and de Vries, D. A., "Moisture Movement in Porous Materials


Under Temperature Gradients," Transactions, American Geophysical Union,
Vol. 38, The American Geographical Union of the National Academy of
Sciences, National Research Council, Washington, D.C., pp. 222, 1957.

Rollins, R. L., Spangler, M. G. and Kirkham, D., "Movement of Soil Moisture


under a Thermal Gradient," Highway Research Board Proceedings, 33, 492-
508, 1954.

Schmill, J. V., "Mathematical Solution to the Problem of the Control of the


Thermal Environment of Buried Cables," AIEE Tran. Vol. 76, p. 175, 1960.

Schmill, J. V., "Variable Soil Thermal Resistivity Steady State Analysis,"


AIEE Trans. Power Apparatus and Systems, V. 86, n2, p. 215-223,
February, 1967.

Shannon, W. L. and Wells, W. A., "Tests for Thermal Diffusivity of Granular


Materials," Proceedings of ASTM, Vol. 47, pp. 1044-1053, 1947.

Slaninka, P., "The External Thermal Resistance of HV Cable in Soil as Non-


Linear Environment," Int. Conf. on Large HV Electric Systems, Paper
Number 21-01-1974 Session.

Stolpe, J., "Soil Thermal Resistivity Measured Simply and Accurately," IEEE
Transactions on Power Apparatus and Systems, Vol. PAS-89, No. 2,
February, 1970.

Taylor, R. L., "HEAT, A finite element computer program for heat-conduction


analysis," Report 75-1, prepared for Civil Eng. Lab., Naval Construction
Battalion Center, Port Hueneme, California, Department of Civil Eng.,
University of California, Berkeley, May, 1975.

Taylor, S. A. and Cary, J. W., "Linear equations for the simultaneous flow
of matter and energy in a continuous soil system," Soil Science of
America, Proc. 28: 167-172, 1964.

Taylor, S. A. and Cavazza, L., "The movement of soil moisture in response to


temperature gradients," Soil Science Society of America, Proc. 18: 351-
358, 1954.

Winterkorn, H. F., "Behavior of Moist Soils in a Thermal Energy Field,"


Proc. of 9th National Conference on Clay & Clay Minerals at Purdue
University, Vol. 9, pp. 85-103, 1960.

Wiseman, R. J. and Burrel, R. W., "Soil Thermal Characteristics in Relation


to Underground Power Cable," AIIE Committee Report, Transaction of
AIEE, Vol. 79, pp. 792-856, December, 1960.

R-2

You might also like

pFad - Phonifier reborn

Pfad - The Proxy pFad of © 2024 Garber Painting. All rights reserved.

Note: This service is not intended for secure transactions such as banking, social media, email, or purchasing. Use at your own risk. We assume no liability whatsoever for broken pages.


Alternative Proxies:

Alternative Proxy

pFad Proxy

pFad v3 Proxy

pFad v4 Proxy