4 hr .. 6‘93. 9‘. l.f!l¢§i.fi1¢\v-.P|. 1 1‘ :1... .. . . 3.3.1.” . . . . .4”. .43.“. 44.3.4.4 4m at. It! 0044 (.11J_fir.«r4m.m 4.“; . .— 143.74.5- {Iii}! 1:01: . \ 1 ..m. 41 wmfiv... .4... , 4 4.4.14.4 u. ....JW...1H4 .15.... N WE f ‘ . 1:11 ”Y 111110 3 1.....4 (aroHvaH‘wl/1 . .(1 71E, .3 ,4 ”flu? .r.‘v\vr. .11.. 13111-311 11111111111 Mm. . e 11 d _ . Cafla 1%. N. E Q. .1 11; L f 'f RATE-01' 1.11 11.1 111011 11133 ' 3'11 T E’r: ., . 1,; .. , 1 .. . 41.7 ,_ .. 3.! .1.v.9..3:!~; .3 «nu-1:81am“ , 1 . ‘2. vinyl-n1 1.. u} .. . . , . _. .. .r. u . 1 , . fr . , . . .. . Ah :' This is to certify that the thesis entitled STRESS AND TIME EFFECT ON THE CREEP RATE OF POLYCRYSTALLINE ICE presented by SH I OU-SAN KUO has been accepted towards fulfillment of the requirements for PH.D. CIVIL ENGINEERING degree in [Maw Major professor Date 9'21'72 0-7639 ABSTRACT STRESS AND TIME EFFECT ON THE CREEP RATE OF POLYCRYSTALLINE ICE BY Shiou-San Kuo The purpose of this investigation was to provide information useful in predicting the time-dependent deformational behavior of polycrystalline ice. Creep tests were performed on fine-grained poly- crystalline ice under uniaxial and multiaxial stress conditions. Stresses ranged between 0.1586 MNm-2.(23 psi) and 2.74 MNm-Z (400 psi). All tests were performed at a temperature of -4.45°C. Two forms of ice specimens were used; solid, cylindrical samples, and thin-walled, hollow, cylindrical samples.‘ All samples were formed by pouring cold, high purity water into clean natural snow contained in a pre- cooled aluminum mold. Ice grain sizes ranged between 1 and 2 millimeters in diameter. The solid, cylindrical samples were tested in uniaxial compression only. The thin-walled, Shiou-San Kuo hollow, cylindrical samples were tested in uniaxial com- pression as well as under multiaxial stress conditions. Creep curves resulting from uniaxial compressive stresses of 0.686 MNm-2 (100 psi), or greater, exhibited primary, secondary and tertiary stages of creep. A constant creep rate was observed in the tertiary stage of creep after the total strain reached about 10 per cent. This observed creep behavior is explained in terms of the mobilization of grain boundary effects, softening of individual crystals due to accumulated strain on their basal planes, and recrystallization. By defining a quantity, S, as the ratio of soft crystals to the total in the polycrystalline mass, a creep equation was derived which contains three terms; a primary creep term, a term representing terminal creep at constant velocity with grain boundary effects fully develOped, and a third term representing the effect of a changing S value. This equation appears to fit experimental uniaxial creep data to strains of at least 25 per cent for the stress range from 0.686 MNm-2 (100 psi) to 2.74 mm"2 (400 psi). Computations indicate an apparent activation energy of 20 Kcal/mole and a flow volume of 2.71 x 104 cubic angstroms based on the rate process theory. 2 Uniaxial creep tests of stresses below 0.686 MNm- (100 psi) exhibited only primary and secondary stages of Shiou-San Kuo creep. Thus application of the derived creep equation was inappropriate. However, experimental results agree with those of Mellor and Testa (1969). Results from tests on thin-walled, hollow, cylindrical ice samples confirm the applicability of the creep equation to multiaxial stress states when stress and strain are considered in the octahedral sense. It appears that the resulting principal strain-rates are parallel and proportional to the corresponding stress deviators within acceptable experimental limits. STRESS AND TIME EFFECT ON THE CREEP RATE OF POLYCRYSTALLINE ICE BY Shiou-San Kuo A THESIS Submitted to Michigan State University in partial fulfillment of the requirements for the degree of DOCTOR OF PHILOSOPHY Department of Civil Engineering 1972 ACKNOWLEDGMENTS The writer wishes to express his sincere appreciation and gratitude to his major professor, Dr. R. R. Goughnour, Associate Professor of Civil Engineering, for his guidance and encouragement throughout the preparation and investi- gation of this thesis. Special thanks are due to Dr. 0. B. Andersland, Professor of Civil Engineering, for his many worthwhile suggestions and support during the experimental work. Thanks are also due to the other members of the writer's guidance committee; Dr. T. Triffet, Professor of Applied Mechanics, and Dr. H. F. Bennett, Assistant Pro- fessor of Geology. To Mr. Leo Szafranski of the Division of Engineering Research Machine Shep, thanks for his cooperation in fabrication of experimental equipment. In addition to those acknowledged above, the writer wishes to thank the National Science Foundation and the Division of Engineering Research at Michigan State University for the financial assistance which made this research possible. ii TABLE OF CONTENTS ACKNOWLEDGMNTS O O O O C O O O O O O O 0 LIST OF TELES C O O O O O O O O O O O C 0 LIST OF FI GURES O O O O O O O O O O O O O NOTAT I ON S O O O O O O O O O O O O O O O ' Chapter I 0 INTRODUCTION 0 O O O O O O O O O 0 II 0 LITERATURE REVIEW 0 O O O I O O O O 0 General Aspects of Creep- . . . . . . Structure of Ice . . . . . . Specific Gravity and Specific VOlume . . The Glide Plane and the Behavior of Monocrystals. . . . . . . . . . Compressive Strength of Ice . . . . . Elastic Behavior of Ice. . . . . . Rate Process Theory and Its Application . Theoretical and Experimental Observations of Mechanical Properties of Ice . . . III. THEORETICAL ANALYSIS. . . . . . . . . Uniaxial Creep of Polycrystalline Ice . . Extension to Multiaxial Creep. . . . . Plastic Strain Computation. . . . . . Solid, Cylindrical Samples. . . . . Thin-Walled, Hollow, Cylindrical samples 0 O O O O O O O O 0 IV. SAMPLE PREPARATION . . . . . . . . . Solid, Cylindrical Ice Samples . . . . Thin-Walled, Hollow, Cylindrical Ice Samples . . . . . . . . . . . V. APPARATUS AND EXPERIMENTAL PROCEDURES . . . Solid, Cylindrical Ice Samples . . . . Thin-Walled, Hollow, Cylindrical Ice Samples . . . . . . . . . . . iii Page ii vi ix 10 12 13 15 16 19 27 43 43 60 64 64 66 69 69 71 74 75 81 Chapter Page VI. RESULTS AND DISCUSSION . . . . . . . . 89 Solid, Cylindrical Ice Samples . . . . 89 High Stress Level. . . . . . . . 89 Low Stress Level . . . . . . . 103 Thin-Walled, Hollow, Cylindrical Ice Samples . . . . . . . . . . . 105 Constant Axial Stress Creep Test. . . 105 Constant Tensile Creep Test . . . . 106 Constant Biaxial Stress Creep Test . . 107 VII. CONCLUSIONS. . . . . . . . . . . . 150 BIBLIOGRAPHY O O O O O O O O O O C O O O 153 APPENDIX. 0 O O O O O O O O O O O O O O 159 iv LIST OF TABLES Table Page 1. Parameters of Equation (3-25) Evaluated for Best Fit to Series B Tests, T = -4.45°C. . 143 2. Parameters of Equation (3-25) Evaluated for Best Fit to Series B Tests, with C and C Values Fixed . . . . . . .1 . . . 144 3. Results of Tests to Determine Value of Recrystallization Constant K '. A11 _2 Tests at -4.45°C and Stress of 1.71MNm (250 psi). . . . . . . . . . . . 145 4. Values of K ', 85’ K ', C ', C ', C3', and KZ'So for a Range of S resses . . . . . 146 5. Value of S as a Function of Stress and Time . 147 6. Parameters of Equation (3-25) Evaluated for Best Fit to Sample 1H. . . . . . . . 148 7. Octahedral Parameters Evaluated for Best_ Fit to Sample 3H Compared with Series B Results . . . . . . . . . . . . 149 Figure 2-1. 2-4. 2-5. 2-6. 3-1 0 3-2. 3-3. 4-1. 5-6. LIST OF FIGURES Schematic RepreSentation of a Typical Creep curve C O O O O O I O C O O O 0 Structure of Ice Lattice (After Pounder, 1961) Stress, Shear Stress and Glide Plane. Temperature Dependence of Young's Modulus for Polycrystalline Ice (After Gold, 1958) . . Schematic of Units of Flow . . . . . . . Schematic Representation of Energy Barrier. . Schematic Diagram of Creep Curve Characterized as Two Parts. . . . . . . . . . . Stress on Ice Crystals in Polycrystalline Mass . . . . . . . . . . . . . Schematic Diagram of Two Phase Loading . . . Disassembled Mold for Forming Thin-Walled, Hollow, Cylindrical Ice Samples . . . . Triaxial Cell, Test Sample, and Mold for Solid, Cylindrical Ice Samples. . . . . Schematic Diagram of Axial Loading System . . Triaxial Cell in Coolant, Text Machine, Cooling System and Sanborn Recorder . . . Deformed Solid, Cylindrical Ice Sample Compared with an Undeformed Sample . . . Schematic Diagram of Triaxial Cell Modified for Testing of Thin-Walled, Hollow, Cylindrical Ice Samples . . . . . . . Nitrogen Cylinder and Testing Apparatus. . . vi Page 11 13 18 20 22 45 45 55 71 75 76 78 82 83 86 Figure Page 5-7. Deformed Thin-Walled, Hollow, Cylindrical Sample Compared with an Undeformed Sample . 88 6-1. Creep Curves for Five Series A Samples . . . 111 6-2. Axial Strain-Rate Versus Axial Strain, Sample 4B. . . . . . . . . . . . 112 6-3. Creep Curve for Sample 6A (a) Showing Three Loading Cycles (b) Loading Cycles Superimposed . . . . 113 6-4. Creep Curve for Several Series B Samples (a) Sample 1B, (b) Sample ZB, (c) Sample 3B, (d) Samples 4B and 5B, (e) Samples 6B, 7B, BB and 9B . . . . . 114 6-5. Creep Curve for Sample 8B Showing Correlation With Equation (3-25) 0 o o o o o a o 119 6-6. Best Fit Values of C as a Function of Stress. 120 1 6-7. Best Fit Values of C from Equation (3-25) Versus Axial Stress . . . . . . . . 121 6-8. Best Fit Values of n from Equation (3-25) Versus Axial Stress . . . . . . . . 122 6-9. Best Fit Values of m from Equation (3-25) Versus Axial Stress . . . . . . . . 123 6-10. Best Fit Values of C from Equation (3-25) Versus Axial Stre s . . . . . . . . 124 6-11. BeSt Fit Values of C from Equation (3-25) -Versus Axial Stre s . . . . . . . . 125 6-12. Axial Strain Necessary for Softening Term to Reach Stated Percent of its Original Value as a Function of Stress . . . . 126 6-13. Creep Curves for Sample 10B Showing Effect of Abruptly Increasing Load (a) First Loading Phase (b) Second Loading Phase . . . . . . 127 6-14. Creep Curves for Sample 118 Showing Effect of Abruptly Decreasing Load (a) First Loading Phase (b) Second Loading Phase. . . . . . . 128 vii 6-27 0 6-28. Creep Curves for Sample 12B, 13B, 14B and 153 in Second Phase Loading . . . . . . . Values of K2' Versus Axial Stress. . . . . Plot of Equilibrium Values of S Versus Axial stress. 0 O O O O O O O O O O O Computed KZ'SS, C2' and K2' So Versus Axial stress. 0 O O O O O O O O O I O Creep Curve for Polycrystalline Sample 16B. . Axial Strain-Rate Versus Axial Stress (After Mellor and Testa, 1969). . . . . Creep Curve for Sample 1H . . . . . . . Creep Curve for Sample 2H (Tensile Creep Test) Creep Curve for Sample 3H . . . . . . . Octahedral Shear Strain Versus Time for Sample 3H. . . . . . . . . . . . Plot of noct Versus Octahedral Shear Stress for Samples 1B through 9B Compared with Sample 3H. . . . . . . . . .. . . Plot of moct Versus Octahedral Shear Stress for Samples 1B through 93 Compared with sample 3H. 0 O O O O O O I I O 0 Plot of C oct. Versus Octahedral Shear Stress for Samples lB through 9B Compared with sample 3H. 0 O O O O O O O O O 0 Schematic Representation of Vectors of Stress Deviator and Strain-Rate for Sample 3H . . viii Page 129 130 131 132 133 134 135 136 137 138 139 140 141 142 O 3’ 3’0 AF GS AH sun 0 ”U NOTATIONS angstrom units = 10"8 cm effective area represented by one flow unit elastic deformation initial inner diameter of hollow cylinder inner diameter change in hollow cylinder» Young's Modulus, energy provided by applied shear stress free activation energy system error grain size heat of activation Planck's constant = 6.624 x 10.27 erg. sec.‘ Boltzmann's constant = 1.3805 x 10-16 erg/°C specific rate of the process net relative distance moved by one flow unit in one transition Avogadro's number = 6.02 x 1023/g mole; number of dislocation hydrostatic pressure free activation energy stress concentration factor universal gas constant = 1.987 cal/mole/°C ix 85 AS St oct. Yoct. Yoct. radius of thin-walled, hollow, cylindrical sample ice density a structure term equilibrium value of S entrOpy of activation tensile hoop stress temperature in °C or absolute temperature time; wall thickness of thin-walled, hollow, cylindrical sample initial volume current volume true axial strain initial elastic strain rupture strain first principal strain, primary creep strain second principal strain, creep strain with grain boundary effects fully developed third principal strain axial strain rate conventional strain shear stress octahedral shear stress shear strain shear strain rate octahedral shear strain octahedral shear strain rate distance between equilibrium positions of flow units distance between flow units normal to direction of flow normal frequency coefficient of viscosity angle between projection of direction onto the glide plane and nearest direction of glide angle between 0 direction and glide plane axial stress first principal stress second principal stress third principal stress stress deviator xi CHAPTER I INTRODUCTION The most highly develOped branch of ice study, and at the same time the only branch that has long been an independent scientific discipline is the study of glaciers, which is called glaciology by West European tradition. But in Latin-Greek the word "glaciology" means literally "the study of ice,‘ and consequently, the study of ice in general, not merely glaciers. Ice structure, as an object of study in structural glaciology, should be understood in the broad sense of the word, beginning with the so-called fine structure of crystals, i.e., the structure of the Space lattice of ice, including structure and texture in the petrographic sense, and going as far as tectonic structures. Pure ice as a crystalline substance is an abstraction. The properties of natural ice bodies are due to the presence of other substances in them, and to the structure of the ice as an aggregate of crystals of a particular form, dimension and mutual orientation. These features in turn are a function of the formation process of the given ice body. Therefore, the physical study of the fine structure and of the constants of the ice mineral is only a beginning, the main requirement being the structural genetic study of ice as a rock. The recent increased activity in the Arctic and Antarctic regions, where ice is used as a construction material, has caused an interest in the mechanical behavior of ice. This interest had previously been linked to the subject of glaciers, and it was long thought the flow of glaciers was associated with regelation. In order to understand the flow of glaciers and ice bodies it is necessary to know at what rate ice deforms under various stresses. Thus considerable work has been done on the deformation of both single crystal and polycrystalline ice, including observation of glacier flow as well as laboratory experiments. The failure of conventional cencepts to describe the time dependent deformation of single crystal and polycrystalline ice has forced many investigators to examine the microsc0pic aspects of creep behavior. Deformation of single crystal ice seems to be controlled by the movement of dislocations on the basal plane. However, the flow of polycrystalline ice cannot be explained so simply. The purpose of this study is to contribute knowledge toward the understanding of the mechanical behavior of polycrystalline ice subjected to an applied stress. Several creep theories have been proposed by many investigators on the basis of experimental findings. The most generally accepted concept is that creep behavior is strongly temperature dependent and is governed by thermally activated processes. Following Kanter's (1938) proposal, theoretical work on creep was influenced by the rate- process theory (Glasstone gt_al., 1941) which was employed to define steady-state creep in metals, and was latter applied to describe the steady-state creep of a large number of materials including ice (Glen, 1954; Steinemann, 1954; Butkovich and Landauer, 1960; Dillon and Andersland, 1966; Goughnour and Andersland, 1968, etc.). Goughnour and Andersland (1968) further suggest that the creep behavior of polycrystalline ice in uniaxial compression is dependent in part on absorbed energy. When a relatively high stress is applied to poly- crystalline ice, after an immediate elastic deformation, the creep curve exhibits a primary or transient creep stage, a secondary or steady-state creep stage, and a tertiary or accelerating creep stage. Typica1.creep curves for this type of performance are shown on Figure 6-4. However, this entire sequence is not always observed during creep of polycrystalline ice. Whether all creep stages are exhibited in any one test depends on temperature, stress, and the duration of the test. A typical creep curve for lower stress (0.1586 MNm-2 = 23 psi) is shown on Figure 6-19, which exhibits only primary and secondary stages of creep. These phenomenological aspects of creep for polycrystalline ice at high and low stress are dis- cussed in the chapters on literature review and experimental results. All the samples for this study were prepared by freezing a mixture of natural snow and precooled pure water. These techniques are described more fully in later chapters. Experimental data obtained from this study indicate that the creep rate of polycrystalline ice indeed depends on the stress, strain, history of loading and temperature. The results of uniaxial compressive creep tests under a certain range of high stresses suggest that the creep rate may approach some constant value at large strains when grain boundary effects are fully develOped and the structure of the ice approaches some equilibrium value. This equilibrium structure value is believed to result from a balance between the creation of soft crystals, created by deformation along their basal planes, and recrystallization. By using this concept along with equation (2-34) developed by Goughnour and Andersland (1968), a creep equation for higher stresses on polycrystalline ice is derived. Data from constant "axial stress creep tests were used to evaluate the param- eters in this equation which appears to account for the creep behavior of polycrystalline ice throughout the primary, secondary and tertiary stages of creep for the range of stresses and temperature (-4.45°C) studied. Uniaxial compressive creep tests at low stresses on solid, cylindrical polycrystalline ice samples were also performed. The constant strain-rate portion of these tests agree very well with the results presented by Mellor and Testa (1969). However, it is believed that the creep equation derived from this study is applicable to low stress tests also, if the duration of creep is long enough. Thin-walled, hollow, cylindrical polycrystalline ice samples were also tested under biaxial stress condi- tions. These results were used to evaluate the validity of the assumption (Nye, 1953), that for polycrystalline ice the principal strain rates are parallel and proportional to the corresponding principal stress deviators. The applicability of the creep equation derived in this study to multi-axial stress states is also verified. CHAPTER II LITERATURE REVIEW Experimental investigations into the creep behavior of ice have followed two main lines of inquiry; behavior of ice single-crystals, and behavior of polycrystalline ice. Several theories have been presented to describe the creep behavior of each. However, only limited success has been attained in correlating the behavior of single ice crystals to that of polycrystalline ice. Chapter II will present a brief review of current concepts of time-dependent deformational behavior of materials in general, and a more specific discussion regarding the creep behavior of both single and poly- crystalline ice. General Aspects of Creep A specimen undergoing continuous deformation under a constant load or stress is said to creep. This phenome- non is observed both in metals and nonmetals. In general, creep refers to strain that is time-dependent and may include elastic, viscous and plastic deformations. From a macroscopic View point, creep deformation is often treated as plastic deformation, and elastic strain and viscous flow are neglected. Creep of metals as well as nonmetals can be demon- strated directly by a creep curve which represents graphi- cally the function between creep strain and time. Its slope at a specific time is called creep-rate or strain- rate. A typical creep curve is shown in Figure 2-1. Tertiary I I Secondary _._IL Strain rr b-—— —-————o — — Time Figure 2-1.--Schematic Representation of a Typical Creep Curve. The shape of the creep curve as shown in Figure 2-1 has led to a subdivision of the creep curve into four parts: l. Instantaneous strain, 80, which is obtained immediately upon loading and exhibits character- istics cf plastic deformation, but includes elastic deformation. 2. Primary creep which is between 80 and £1. The creep rate decreases continually. 3. Secondary or steady-state creep which is between 81 and £2. The creep rate remains constant, indicating a nearly steady-state condition. 4. Tertiary creep which is beyond 6 The creep 2. rate increases continually, leading eventually to rupture at strain Er and time tr' This behavior is thought to be a result of inter- crystalline movement as opposed to the intra- crystalline movement of primary and secondary creep. Tertiary creep can occur at any temperature providing the stress level is relatively high. Whether all creep stages are observed in any one test depends on stress, temperature and the duration of the test. Different materials will show similar creep curves I under proper adjustment of stress and temperature. Creep curves for metals, crystalline nonmetals and noncrystalline materials are all similar. Considering the differences in atomic structure of the materials, it must be evident that the details of the creep curve itself cannot identify the micro-mechanisms responsible for creep. However, the similarity of the creep curves for these materials does indicate that a similar sequence of rate determining mechanisms is followed in all cases. Experimental evidence conclusively shows that creep flow is thermally activated; local thermal agitation pro- vides added energy, beyond that provided mechanically, to overcome barriers to creep deformation. Then the applied stress aids in overcoming these barriers but serves also to give direction to the resultant flow. A generalized function for the creep rate, 6, based on the theory of thermodynamic fluctuation can be given by -AFi(T,S) é = 1 z1 (v,T,S)oi (T,S)exp [-——§§————] (2-1) The function, 21' may depend on the temperature,T, and contains the normal frequency, v, for the flow unit, the entropy change and a structure term S. The latter involves such microstructural features as grain size, precipitation and dispersion of one or more phases, density, distribution, and length of dislocation. The stress function oi may depend on temperature and structure because creep mechanisms are temperature- and structure-dependent. AFi governs the creep rate and is temperature and structure-dependent. 10 R is the universal gas constant. From equation (2-1) it is clear that creep is a Complex phenomenon, and that the evaluation of this function requires extensive experimental and theoretical work (Frank Garofalo, 1965). Structure of Ice The prOperties of ice, as is the case with all crystalline materials, are determined not only by its chemical composition, but also by the geometry of its space lattice and the nature of the bond that keeps the structural units of its space lattice in equilibrium. The problem of the fine structure of ordinary ice has not yet been solved completely. X-ray analysis can establish only the position of the oxygen atoms in the space lattice of the ice. The hydrogen atoms do not scatter x-rays sufficiently, and at present their position is judged primarily from indirect data. The position of hydrogen atoms changes continuously at temperatures above approximately -70°C, so that only the average.structure determined by the position of the oxygen atoms can be recorded. The unit cell of this average structure consists of four oxygen atoms. These oxygen atoms are formed in a tetrahedral pattern, with each oxygen atom being surrounded by four approximately equally spaced oxygen atoms at the vertices of the tetrahedron. Each pair of oxygen atoms 11 is linked by a hydrogen bond (Pounder, 1965). The unit cell of ice is depicted on Figure 2-2. 0'" OI" Olll 0.923 A 2.760 A Figure 2-2.--Structure of Ice Lattice. (After Pounder, 1961). If the tetrahedron were perfect, that is if all bounds were of equal length, and all the 0'00" and O"OO" angles were equal, these interior angles of the tetra- hedron would be 109° 28'. Solid lines shown on Figure 2—2 represent hydrogen bonds and the outline of the tetra- hedron surrounding the oxygen atom marked 0 is shown by dashed lines. The three atoms marked 0" form an equi- lateral triangle in a plane which is called the basal plane of the lattice. The 0'0 bond gives the direction 12 I known as the’C-axis, which is perpendicular to the basal plane. Considering now the four bonds of the atom 0', one is O'C, and the other three are O'O"'. These three atoms must also lie in a plane perpendicular to 0'0. Thus the tetrahedron about 0' is inverted, with its base parallel to that of the tetrahedron about 0. Figure 2-2 shows the complete bonding system for the O and 0' atoms, but each of the three 0" and 0"' atoms must have other three additional bonds. In this way the lattice is extended in all three directions. This orderly arrangement of atoms is disrupted at grain boundaries. Masses of polycrystalline ice are aggregates of single crystalline grains. In quickly frozen masses these grains may range from only a fraction of a millimeter to several centimeters in size. The boundaries between neighboring crystals in polycrystalline ice are usually irregular and often depend on the mode of freezing. Specific Gravity and Specific Volume Ice has an "open," not very compact structure. Therefore it is lighter than water. The specific gravity of pure ice at 0°C and 1 atm pressure is 0.9168 gram/cm3, and its specific volume is 1.0908 cm3/gram (Bridgman, 1912; Lonsdale, 1950), while the specific gravity and specific volume of water are 0.999863 gram/cm3 and 1.000132 cm3/gram, respectively. Any departure from the above 13 value for ice indicates the presence of impurities, or a different temperature. The Glide Plane and the Behavior of'Monocrystals The basal plane of an ice crystal is its well defined glide plane. The stress, 0, at the elastic limit differs from the critical tangential stress I as a function of the angle ¢ between the c direction and the glide plane, and also as a function of the angle w between the projection of the c direction onto the glide plane and the nearest direction of glide:(see Figure 2-3) I = o(sin ¢ cos 0) (2-2) L\‘<::z"glide plane Figure 2-3.--Stress, Shear Stress and Glide Plane. In ice the magnitude of the critical normal stress p necessary to rupture the crystal along the basal plane is incomparably greater than the critical tangential stress 14 at which gliding starts. If the critical tangential stress I is reached, the ice will deform plastically, without disruption of continuity; if, however, the criti- cal normal stress p is reached at the same time or subse- quently during the plastic deformation process, the ice will act as a brittle body and split. A force causing shear may act in three main, mutually perpendicular directions (McConnel, 1890); 1., The shear force coincides with the basal ‘plane, only translation occurs, the deforma- tion is plastic; Shear force IIIII lllll Main axis 2. The shear force acts in the direction of main axis perpendicular to the basal plane. The elementary plates of the crystal bend and break after the critical stress is exceeded. The deformation is elastic-plastic; 15 Shear force "lI-IIIIIII" . v Main axis 3. The force acts in the basal plane, but the shear force is perpendicular to it. The elementary plates that are rigid in the longi- tudinal direction allowonly a negligible elastic deformation, which changes to disrup- tion in the case of a significant increase of stress. Generally there is no plastic deforma- tion in this case. Compressive Strength of Ice The compressive strength of ice has been extensively investigated. An excellent summary of the results prior to 1951 may be found in the companion volume of U.S. Army Snow,Ice and Permafrost Research Establishment Report Number 8 (Volume II, Appendix B, Table 12). According to the report, Vitman and Shandrikov found a general increase in the strength of ice as the temperature is decreased. A sample having a number of small crystals has a greater strength than a sample having one or more crystals. Butkovich (1954) performed a series of tests on cylindrical ice specimens to determine the ultimate 16 strength of ice. He used lake ice, natural snow ice, and commercial ice. His results showed conclusively that temperature has a pronounced effect on the strength of ice. His results produced a strength variation of as much as 140 psi from the mean strength for ten samples. An important conclusion is that the strength is indeed dependent on how the load is applied to the ice itself. Leonards and Andersland (1960) found that the strength of ice increases approximately 15 psi per degree centigrade decrease in their unconfined compression tests using identical, polycrystalline cylindrical samples. Their results showed good agreement with those of Butkovich (1954). Elastic Behavior of Ice The determination of Young's modulus for ice has been attempted by two methods, static and dynamic. Dorsey (1940) questioned whether significantly useful values can be obtained by means of the static method. There are several factors which could affect the values of Young's modulus: temperature, loading rate, grain size, orientation, and density. Many of the early static determinations of Young's modulus for ice were done at temperatures only slightly below 0°C, and those tests which were performed at tempera- tures below 0°C were done with little concern for the control of temperature. The results from U.S. Army Snow, 17 Ice and Permafrost Research Establishment (1952) show that as temperature decreases, the value of Young's modulus increases slowly. Their linear relationship gives an E value of 1.25 x 106 psi at 32°F and 1.52 x 106 'psi at -34°F. These values were obtained from both sonic and static test methods. Nakaya (1959) observed a hysteresis phenomenon in the relationship between Young's modulus and temperature. In most polycrystalline aggregates the likelihood of large dependence on orientation is small. However, Young's modulus depends on the preferred orientation of the grains (Van Vlack, 1959) and its value from specimenS‘ prepared with random orientations will represent an approxi- mate average value. Gold (1958) performed experiments to determine the elastic modulus of ice by a static method. He took care to control both temperature and grain size. Figure 2-4 shows his results plotted as Young's modulus versus temperature. It is seen from Figure 2-4 that as the grain size increases, the modulus increases. Dynamic measurements of Young's modulus were also cede With little attention given to control of the experi- mental variables. However, at the higher "loading rates" used in the dynamic method, the temperature dependence ‘was a full order of magnitude less than that observed for static meaSurements.f This result is shown on Figure 2-4 1.4x106 1.3 '5: ,3 1.2 m 3 , 1.1 8 5‘ 1.0 P m 5 0.9 O >0 0.8 18 10.0x1010 L E sonic = 9.0 x 1010 to 10 x 1010 dyn/cm2 - Ex =- 8.34 x 101'0 dyn/cm27 7 9'0 l l ——A——oa—%1 44—33—— ‘9 8 as 4 2 “s 8 ”a - .0 O . a 8' I- q 7.0 g '5‘ £19 3 I E: 13.1 12 .. 6.0 -2 10 2 o s .9 _ Ex I (5.69 - 6.58 x 10 T)x10 dyn/cm. a. . 5 1 J n n L 1 J 5.0 g -40 -35 -30 -25 -20 -15 -10 -5 0 Temperature °C LEGEND [3 Single Crystal .Fine Grain (diam. < 0.50 cm) Medium Grain (0.50 < diam. < 1.50 cm) D> D' C) Large Grain (diam. > 1.50 cm) 0 Average Grain diameter not observed Figure 2-4.--Temperature Dependence 6f Young's Modulus for Polycrystalline Ice (After Gold, 1958). 19 also. The dynamic values are seen to be very consistent. Gold (1958) stated that doubling and halving the loading rate did not appreciably change the Young's modulus values. Halvorsen (1959) found that his values of elastic modulus varied between 1.735 x 105 psi and 1.94 x 105 psi as the loading rate varied. The density of ice has a pronounced effect on the value of Young's modulus (Nakaya, 1959; Halvorsen, 1959). As the density decreases, a corresponding decrease in the value of Young's modulus occurs. Nakaya's results show that for densities between 0.915 and 0.903 Young's modulus 6 varies from 1.33 x 10 psi to 1.09 x 106 psi respectively. His results agree well with those of previous investigators. Rate Process Theory and Its Application The reaction-rate theory (Glasstone, Laidler, and Eyring, 1941) which was employed earlier by Eyring (1936) to predict viscosity, plasticity and diffusion in viscous fluids has greatly influenced the theoretical work on creep. The reaction-rate theory involves the concept of activated complexes and is based on statistical mechanics. The formation of activated complexes requires energy. A complex is regarded as being situated at the top of a free energy barrier lying between the initial and the final states. It is assumed that the "unit of shear process" involves the motion of "units of flow" which are regarded as elementary structures within a solid (atoms, molecules or aggregates of molecules). The motion of these flow 20 units passing from one equilibrium position to another constitutes the shear process, but can only be accom- plished by overcoming energy barriers. This barrier can be overcome by activating the flow units." A schematic representation of units of flow is shown on Figure 2-5, where T is the shear stress, 1 is the distance between equilibrium positions of flow units, and Al is the distance l {er tom 1. Figure 2-5.--Schematic of Units of Flow. between flow units normal to the direction of flow. If K' is the net number of transitions per second made in the direction of shear, then the shear rate dy/dt becomes 21% I K'. (2-3) >a rely An expression for the specific reaction rate equation of molecular processes based on the rate process theory is 21 K' = %? e-AF/RT (2-4) where K' is the specific rate of the process, k is Boltzmann's constant equal to 1.3805 x 10"16 erg/°C/mole, T is the absolute temperature, h is Planck's constant equal to 6.624 x 10.2 erg. sec., AF is the free activation energy, and R is the universal gas constant equal to 1.987 cal/degree mole. In the absence of an applied shear stress to the material, a state of equilibrium exists and the number of transitions in one direction is equal to those in the opposite direction. If, however, a directed shear stress is applied to the material, the energy barrier becomes distorted. Kauzmann (1941) assumed that the activation energy for motion of units of flow in the direction of shear is lowered by the applied shear stress and the change in activation energy is proportional to the stress. In the direction opposite to the direction of flow the Tactivation energy is raised by an equal amount. Figure 2-6 depicts the energy barrier schematically. In Figure 2-6 the solid line represents the energy barrier in the absence of applied shear stress, and the dashed line shows the distortion of the barriers when shear stress is applied. The energy, E, provided by the applied shear Stress is given by Potential Energy 22 Shear Stress T ' —* I ' I I ' I I With applied l shear I l Without applied I I, shear I I I I I I \ I ’1’ AF \ I TqA£ —I I I I 2 \_1| \ I 2/2 \\ 14:2 I ~ ,’ £ Direction of Strain Figure 2-6.--Schematic Representation of_Energy Barrier. E = rqu (2-5) where A is the projected area of the unit of flow on the shear plane, 1 is the distance the shear stress acts in surmounting the potential barrier, and q is a stress con- centration factor. This is the energy required to displace a flow unit from its initial state to its final state. 23 As a result of the applied stress on the unit of flow, the barrier height in the forward direction is (AF - l%?£) and in the backward direction is (AF + 1%?£). The net specific rate in the forward direction is then kT AF IqAR kT TgAR) - —- AF I _ I __ __ _ __ __ .- K K 11 9X9 I RT + 2RT) 11 EXP I RT' 2RT 2kT _ A2. . TgAR _H— exp ( RT) Sin h ZRT' (2-6) By substituting this into equation (2-3) the shear strain-V rate becomes -_A___2kT -92: . - A1 h exp ( RT) Sin h TgA2 ZRT‘ (2'7) .< . ll 9:? Herrin and Jones (1963) indicated that A and Al are only approximations of the order of the size of a flow unit, and it may be assumed that A equals Al. Hence equation (2-7) becomes 2kT ' . g l y = -3— exp (-%%) Sin h T A 2RT' (2'8) Since k (Boltzmann's constant) equals R/N, where N is Avogadro's number (6.02 x 1023/9 mole), changing the units of equation (2-8), one obtains °_ 2kT _ £3. . TgA£ _ 7“ h exP I RT) Sin h 2kT' (2 9) 24 From thermodynamics, the free energy of activa- tion AF can be expressed as a function of temperature, AF = AH - TAS ' (2-10) where AH is the heat of activation, and AS is the entrOpy of activation. Substituting equation (2-10) into equation (2-9) gives _ 32_ '-AH + TAS . TgAR _ y — 2 11 exp ( RT ) Sln h 2kT (2 ll) or I = 2%? exp (%§) exp (- %¥) sin h l%%%u (2-12) This is the general creep equation derived by using the rate process theory, and is based on a single molecular creep mechanism. However, this need not always be true since in a given material different mechanisms may be operating simultaneously, each having its own characteristic values of AH, AS, q, A, and 2 (Kauzmann, 1941). Thus .equation (2-12) may be written as AS. AH. (qu).I '_ 152 _.1_ -.__1.- ___1_ - y -i 2 11 exp ( R ) exp ( RT) Sln h 2kT (2 13) where the sum is over all possible mechanisms. Conrad (1961) pointed out that usually only one of the mechanisms (i.e., only one of the terms in equation (2-13)) will 25 account for the greater part of the observed flow rate, but it is possible that a single term will not give a major part of the shear rate under all conditions of stress and temperature. For example, one mechanism may account for most of the creep under low stresses, while under large stresses a different mechanism may dominate. Each mechanism contributes something to the total observed rate under all conditions. When the applied shear stress is small compared to the thermal energy, Tqu TgA£ 31“ h 2kT ~ 2kT. Equation (2-9)becomes ° _ 2kT _ A§_ Tqu Y ‘ h exP I RT) 2kT or .. Tel-ftfiv eXp (_ %). (2'14) .4 I That is the shear strain rate is proportional to the shear stress which constitutes Newtonian flow. The coefficient of viscosity can be expressed as _ h.. A}: .. n — qu exP (RT). (2 15) 26 When the energy supplied by the applied shear stress is much greater than thermal energy, then . TgAl z 1_ TgAl. 31“ h 2Kt 2 ex? 2kT' Equation (2-9) will become ° kT AF 2 'Y = -h— exp (" .1371?) EXP ("l'_2?<_%_) . (2‘16) Equation (2-16) is a very useful form in the study of the creep proCess. Equation (2-12) can also be written to the form y = A Sin h B (2-17) .. M 9.5. _ All where A — h exp (1!) exp ( RT), and _ 2kT B — qAR'o Born and his co-workers (1954) have shown that for poly- crystalline aluminum, B is independent of temperature and stress and appears to be independent of the structure of metal. A is dependent on temperature, structure and stress (Conrad, 1961). Dorn and his co-worker (1954) found the creep rate equation as -AH/RT Bo e e m. = 5p (2-18) 27 Where the stress function eBC is the effect of stress on the deformation process itself and not on the structure, Sp is the effect of structure on the creep rate. Theoretical and Experimental Observations of_MechanicaI Properties of Ice Investigators have studied the flow of ice, as influenced by stress and temperature by observing the deformation of ice single crystals and polycrystalline ice. In 1887 Main performed deformation experiments on ice samples. He showed that polycrystalline ice would flow under conditions that would preclude regelation, but his control of temperature and other variables pre— cluded any quantitative conclusions. McConnel and Kidd (1888) followed this work to confirm Main's investigations. In addition, they found a marked temperature dependence of the flow rate of polycrystalline ice. They did work also on single ice crystals, and found that when the C-axis is perpendicular to the stress, the crystal yields only slightly. However, they felt that ice crystals were not plastic and attributed to the flow of polycrystalline ice to "some action at the interface of the crystal." McConnel (1891) continued the work on single crystals and found that ice crystals are plastic and slide on the basal plane. He likened ice to an "infinite number of thin sheets of paper, normal to the optic axis." If ice 28 deforms more or less perpendicularly to the main axis, the slight elastic stress relaxes intensively, while if the deformations are parallel to the main axis, the relaxation is weaker. Hawkes (1930) found that deformation in ice is due to slip along (1) basal plane, (2) grain boundaries, ‘and (3) fracture surfaces. I In recent years, an increasing number of researchers have investigated the flow of ice. Glen and Perutz (1954), and Steinemann (1954) have shown that ice deforms plastically only by translational gliding on the basal plane (0001), which confirms the original observations of McConnel (1891). Steinemann (1954) found that ice single crystals subjected to shear stress parallel to the basal plane undergo a transition in creep behavior at strains between 10 to 20 per cent. This transition sometimes occurs abruptly and at other times is gradual, but it always results ulti- mately in a drastically increased yet uniform creep rate. The relation between stress, strain and time for ice single crystals has been investigated by several experi- menters. Griggs and Coles (1954) performed creep tests on single crystals for which the basal plane was oriented at 45° to the compression axis. They found that their creep data could be best fitted to an equatiOn of the form a = a[(o - sz) t]2 (2-19) 'where e is the compreSsive strain, 0 is the stress, T is the temperature below the melting point and t is the time. 29 If e is expressed in per cent, a in Kg/cm.2, T in °C and t in hours, then a equals 0.62 and b equals 0.02.' Steinemann (1954) carried out his tests on ice single crystals for more than 20 hours and found that his creep curves appeared to settle down to a steady rate of flow rather than continue to increase parabolically, and he has related steady-state rate by plotting shear strain rate Y against shear stress T logarithmically. He found a relationship of the form § = KTn, where K and n are constants. Butkovich and Landauer (1959) also found that the flow law, § = KTn, provided a best fit to data from shear tests of single crystals in "easy glide." Steinemann found a value of n at low strains ranging from 2.3 to 3.9, and at higher strains of 1.5. Griggs and Coles (1954) indicated a stress-dependence of the flow rate corresponding to an n value of 2.0. Butkovich and Landauer (1959) found n average of 1.7. Kamb (1961) performed torsion tests of hollow single crystal ice cylinders and deformed the specimen to several hundred per cent at a temperature -2.9°C. He found a value of n equal to 2.5. Jellinek and Brill (1956) applied a different approach. They performed tension creep tests on single crystals with their optic axes oriented at 45° to the sample axes. They found that some of the samples followed a sat2 law, whereas others showed a linear relation between strain and time. They assumed first that 30 e = KlN (2-20) where N is the number of dislocation in the specimen at time t, and then u = K or N = KZt + NO (2-21) where No was the number of dislocation present at the start of the creep test. From equations (2-21) and (2-20) one obtains a = Klet + KlNO. (2-22) Integrating equation (2-22) gives H _ _ ' 2 _ e — 2 Klet + KlNot + so (2 23) where so is the instantaneous elastic strain. If the Second term is negligible, a parabola is obtained. However, if No is quite large, then the linear term might dominate and a linear relationship results. As the number of dis- locations increases with time, they will mutually annihilate each other and finally reach steady-state creep. This was also observed by Steinemann (1954) as mentioned previously. 31 Higashi and Sakai (1952) performed experiments on. the movement of low-angle tilt boundaries in ice and found an activation energy between 12 and 17 Kcal/mole. Itagaki (1966) found that the activation energy for self-diffusion in ice varied from 12.6 Kcal/mole to 15.4 Kcal/mole. In polycrystalline ice, where the orientations of individual grains are random, continuity demands that mechanisms other than basal slip operate to permit the grains to conform to an arbitrary change of shape. Glen (1952) reported on the results of compressive creep tests on blocks of polycrystalline ice at -l.5°C. His results showed that polycrystalline ice does not possess a con- stant coefficient of viscosity, but obeys a relationship between compressive stress, a, and strain rate, é, of the form é = kcn (2-24) where n is a conStant equal to about 4. Later in 1955, Glen continued his experiments on the creep of poly- crystalline ice under compressive stress conditions and found that the relationship between the minimum observed creep rate, s, the applied stress, 0, and the absolute temperature, T, can be represented by s = B exp (- %%)on (2-25) 32 where R is the gas constant, and B, n and Q are all constants. He found n equals about 3.2, Q equals 32 Kcal/ 24 if the stress ismeasured in mole, and B equals 7 x 10 bars and the strain rate in years-1. The activation energy, Q, equal to 32 Kcal/mole found by Glen is higher than that found by other investigators. Landauer (1957) pointed out that this value is strongly dependent upon the use of one creep Specimen at -12.8°C, and a liberal error could be present in the value. Landauer (1957) did not disprove Glen's value of Q, as he was attempting to relate the activation energy of snow to ice. For snow, Landauer found the activation energy varied from 14 Kcal/mole to 27 Kcal/mole. Glen (1955) observed that polycrystalline ice seems to follow an Andrade type of behavior in the transient creep region if one ignores acceleration due to recrystallization. From Andrade's law 2 = 2 (1 + Bt1/3)eKt 0 (2-26) where 2 is current length, £0 is original length, 8 and K are constants, and t is time. In terms of the logarithmic strain, this relation- ship becomes a = 1n (1 + Btl/3) + Kt (2-27) 33 . 1/3 . or if 8t << 1, equation (2-27) becomes a = Btl/3 + Kt (2-28) Glen assumed that the transient portion followed a tl/3 relation. If it was subtracted, then the straight line became K = 0.01704"2 which showed a stress exponent of 4.2. Glen (1955) stated that when polycrystalline ice is sub- jected to creep, its grains may be either slipping rigidly past each other, with consequent boundary changes (grain boundary creep) or else accumulating strain inside the grains. In the latter case, the probability of recrystal- lization would rise as the internal strain increased, and new recrystallized material would be continually appearing. After a long time a steady-state would be reached in which the hardening caused by the rise in internal strain is balanced by the softening caused by recrystallization. A simple dimensional argument could be used to relate this strain to the grain size, if the functional relationship of both nucleation and growth rate were known as a function of strain. The lack of an accelerated stage in creep curves at lower stresses is due to the fact that these samples had not accumulated as much total strain as the others, and would have developed acceleration later if the duration of test was long enough, or it might be due to recrystallization occurring too slowly for the final steady-state to be reached. 34 Steinemann (1954) wrote the power law equation in the form § = KTn, where I is shear strain, K and n are constants, and T is shear stress. Using both compressive and tensile creep measurements on randomly oriented, small- grained ice crystals, he found a value of stress exponent n equal to 4.16 at high stresses. The value of n decreased with decreasing stress. He pointed out that a simple power law did not adequately fit all of the data. However, he observed recrystallization during the measurements, similar to that postulated by Glen (1955). He stated that recrystallization takes place when crystals are deformed inhomogeneously above a certain amount. In poly- crystalline ice, recrystallization takes place during the experiment if the applied stress is sufficiently high, or if the time of testing is long enough for the strain to attain the necessary value. Thus, polycrystalline ice remains hard by recrystallization. If this recrystalliza- tion were absent, a marked transition followed by a linear part could be expected. Rigsby (1961) also found that under large shearing stresses ice recrystallizes in such a manner that the newly formed crystals are favorably oriented for basal slip. Nye (1953) found a value of 3.07 for the stress exponent from tunnel-closing experiments in glaciers. Butkovich and Landauer (1958), measuring shear creep on various types of polycrystalline ice, found that within a 35 broad band, all the data fell on a straight line when plotted as log é versus a with a stress exponent equal to 2.96. At lower stress levels Jellinek and Brill (1956), performing compressive creep tests on randomly oriented polycrystalline ice with noticeable porosity, found a stress exponent of unity and activation energy equal to 16.1 Kcal/mole. Butkovich and Landauer (1960), also performing compressive creep tests on randomly oriented ice with noticeable porosity (average density 0.905 gram/cm3) and on large grain columnar ice with no porosity, obtained values between 0.86 and 1.15 for the stress exponent, and a value of 14.3 Kcal/mole for activation energy. Bromer and Kingery (1968) performed tensile creep tests with temperatures from -3° to -13°C at stresses between 2 x 105 dyne/cm2 (2.9 psi) and 2 x 106 2 dyne/cm (29 psi) on millimeter range grain-size polycrystalline ice. They analyzed their data by using five theoretical models: (1) grain-boundary sliding by Kauzmann (1941), (2) grain- boundary migration by a diffusional process involving atomic or molecular transfer, (3) microcreep as described lay Chalmers (1936), Cottrell and Jaswan (1949), and Friedel (1964), (4) diffusional creep as described by Nabarro (1948) and Herring (1950), and (5) diffusional creep with grain-boundary diffusion rate controlling as described by Coble (1963). They developed the following equation: 36 m ll Ac (ss)-m (2-29) or —(1> m ' n - 3: (GS) (2'30) where s is strain rate, 0 is stress, GS is the grain-size, n is the viscosity, and A and m are constants. They found that the Nabarro-Herring model provided the best fit for their data. Their results showed that polycrystalline ice behaves in a viscous manner, i.e., the strain-rate is proportional to the grain size squared. The flow process has an apparent activation energy of 12 Kcal/mole. Mellor and Testa (1969) following Bromer and Kingery (1968), performed uniaxial compressive creep tests on fine-grained polycrystalline ice at a temperature of -2°C and a stress range between 0.1 and 0.5 Kgf/cm2 (1.45 psi and 7.25 psi). (Their tests were carried for a total duration of 10.5 months. They showed that the stress exponent of a simple power relation between strain-rate and stress was 1.8. Therefore, the earlier results by Bromer and Kingery (1968) suggesting linear viscous behavior at low stress were believed to be invalid. Their experiment was simple but the results were excellent. The stress exponent and activation energy obtained at lower stress levels are quite different from those obtained at higher Stress levels. It appears that there must be more than a single mechanism reSponsible for 37 plastic flow in polycrystalline ice. Glen (1955) has ‘pointed out that there might be a transition from inter- granular to intragranular deformation. Butkovich and Landauer (1960) also indicated that there might be two mechanisms operating when they proposed a flow law having a linear and a cubic term in the form . I 3 _ Y = at + b1 (2-31) where 4 is shear strain-rate, T is shear stress, a and b are constants. Dillon and Andersland (1966) in their com- pressive creep tests on polycrystalline ice concluded that a transition from one mode of creep behavior to another appears to separate the high and low stress regions. As mentioned previously, by theoretical considera- tions, temperature has a strong influence on the creep rate of polycrystalline ice. Most theories arerased on the rate process theory. Jellinek (1960) performed experi- ments which imposed plastic deformation of thick-walled snow-ice cylinders under hydrostatic pressure. His idea basically followed the model presented by Kauzmann (1941). Dillon and Andersland (1966), applying the rate process theory, analyzed their results on the creep of cylindrical polycrystalline ice and proposed an hyperbolic sine stress and steady-state strain-rate relationship. They proposed an equation in the form 38 ' T Yoct. = Yo sin h -%SEL (2-32) 0 _ where Ioct. and Toot. are respectively the octahedral shear strain rate and shear stress, and (o and T0 are constants. At low stresses, values of (o vary slightly with temperature. However, the creep curves shown in their paper had a maximum duration of only 80 minutes and a total axial strain of 2 per cent. This may be too short a period or too small a total strain to establish the minimum creep rate in their range of stresses and tempera- tures. Goughnour and Andersland (1968), using experimental apparatus similar to that used by Dillon and Andersland (1966), examined the creep oprolycrystalline ice through the use of compressive tests at high stress. They found that the creep rate of polycrystalline ice depends on the stress, temperature, strain and the absorbed strain energy. Based on their data they proposed creep equation of the form to II K1 exp (-nls) + K2 exp (n2 f ode) (2-33) and (0 ll K1 exp (-nlt) + K2 exp (nzt) (2-34) 39 where e is the axial plastic strain, 0 is the axial stress difference, t is the time, and K1, K2, n1 and n2 are conStants that depend on stress and temperature. In these equations they assumed that the total effect of all mechanisms contributing to a decreasing creep-rate could be represented by one term which they called the hardening term, the first term in both equations, and that the total effect of all mechanisms causing an increasing creep-rate could be represented by the second term which they called the softening term. These equations appeared to describe the axial creep behavior of polycrystalline ice throughout primary, secondary and tertiary creep for strains up to about 5 per cent and the ranges of stress and temperature they considered. From the results they indicated that equation (2-34) fitted their data better than equation (2-33). Their creep tests were carried to strains of only 5 or 6 per cent. This strain level may again be too small to adequately demonstrate tertiary creep behavior. During the course of tests to investigate crack deformation in polycrystalline, columnar grained ice sub- jected to a constant compressive loading, Gold (1963) observed several deformation mechanisms. Some of these have been described by earlier investigators. The deforma- tion mechanisms as described by Gold are: 40 1. Slip Bands. Near the melting points single crystals with hexagonal symmetry usually deform by slip (on the basal plane. The slip lines define the trace of the basal plane at the ice surface. 2. Grain-Boundary Migration. Grain-boundary migration was one of the first signs of change in the grain-boundary configuration. Every boundary observed showed evidence of migration. Gold felt that grain- boundary migration was necessary to relieve the constraints imposed by the grain geometry. 3. Kink Bands. Kink bands formation is a mechanism by which a bending moment, transverSe to the slip direction, can be relieved in crystals with only one or two possible slip directions. 4. Distortion of Grain Boundaries. As the deforma- tionincreased, the boundaries of most grains became severely distorted. 5. Crack Formation. Crack formation is time dependent as well as stress dependent. The formation of the crack may be aided by the growth under the applied Stress of potential crack nucleation sites such as cavities. Once the crack has formed, the grains associated with it can more easily conform to the applied deformation. For creep under severe deformation these can be considered as "accommodation" cracks. 41 6. Cavities. The cavities were observed in the region of grain boundaries, grain-boundary triple points, and the intersection of slip planes and subboundaries. }Sometimes cavity formation was great enough to form a con- tinuous column. There was evidence that cavities could serve as nuclei from.which cracks could develop. 7. Recovery and Recrystallization. 8. Unusual Creep Behavior. Gold found that ice under compressive creep sometimes exhibited an initial increasing creep rate. This is unusual creep behavior. The deformation mechanisms observed by Gold are in general agreement with the scheme proposed by Shoumsky (1958). Gold's experimental results suggest that for a given temperature, creep rate is probably the major factor in determining what mechanisms are required for a grain to conform to the applied deformation. At low creep rates grain-boundary migration, subboundary, Kink band, and Cavity formation, polygonization, and bending are likely to be sufficient. For higher creep rates accommoda- tion cracking during the primary stage and greater dis- tortion in the grain boundary region occurs. In addition to creep rate, the amount of deformation and the duration ofloading are probably important also. It appears that the deformation of simple crystals proceeds primarily by means of basal slip; however, evidence of nonbasal slip has been observed as described 42 earlier. Polycrystalline ice presents a much more compli- cated picture. At low stresses ice may behave in a quasi- viscous manner. At high stresses a different mechanism seems to dominate. Several suggestions and proposals have been made to account for the flow of polycrystalline ice. At the present time, none of the flow laws seems entirely satisfactory throughout the primary, secondary and tertiary stages of creep. As evidenced by the onset of tertiary creep, extension of the creep to higher deforma- tions leads to recrystallization and an apparent softening or structure change of the ice sample. CHAPTER III THEORETICAL ANALYSIS Development of theory to explain the time dependent deformational behavior of polycrystalline ice, based on the observed behavior of ice single crystals will be pre- sented in Chapter III. The theory will first be developed for uniaxial deformation and then extended to include multiaxial creep states. The experimental portion of this study included uniaxial creep tests on solid, cylindrical samples, and multiaxial creep tests on thin-walled, hollow, cylindrical samples. Equations for computing the plastic strains for these samples are also derived. Uniaxial Creep of Polycrystalline Ice It appears from the previous discussions that there are two flow regimes for polycrystalline ice. One appears to dominate at high stresses and it exhibits three stages of creep. Thus far, it is not clear how the flow rate depends on the stress under these conditions. The second mechanism becomes important at lower stresses and will be discussed with experimental result which may account for the observed flow behavior in polycrystalline ice. 43 44 In order to develop an equation which will predict the plastic behavior of ice it is necessary to work with data that represents plastic deformation only. The theories presented in this study deal only with plastic deformation. Creep such as that illustrated on Figures 6-1 and 6-4, may be characterized as consisting of two parts. The first part 61, limited to some fixed value, represents primary creep and may be due to the increasing importance of grain boundary effects. The second part 62, represents creep with grain boundary effects fully developed, and the magnitude of the strain-rate e for a given stress would 2: depend primarily on the structure of the crystals. A schematic representation of these mechanisms is shown on Figure 3-1. Figure 3-2 is used to explain these mechanisms in a microscopic sense. The polycrystalline mass shown is initially relatively stress free. When shear stress is applied, grain A is oriented for easy glide, while grain B is not. Consequently as grain A deforms stress concentra- tions deve10p at the grain boundary in the area indicated, and a larger portion of the applied shear stress is trans- ferred to grain B. Since grain B is not oriented for easy glide the overall creep rate of the polycrystalline mass decreases. For deformation to continue, mechanisms such as those described by Gold (1963) are necessary to 45 Figure 3-l.--Schematic Diagram of Creep Curve Characterized as Two Parts Basal plane orientation Stress concentrations in this area Grain B \ ‘1' \ 4——— Applied shear stress Grain A Figure 3-2.--Stress on Ice Crystals in Polycrystalline Mass. 46 relieve these stress concentrations. One of these mechanisms is recrystallization. Recrystallization will affect softening since the newly recrystallized material is initially stress free. This is essentially the same mechanism as described by Gold (1963). The time law of creep during the course of deforma- tion as demonstrated above may be expressed as (3-1) Recall equation (2-12), the general creep equation derived by using the rate process theory, .=fl é—S_ —é—Ii ' TAR «- Y 2h exp (It) exp ( RT) Sln h —%kT (3 2) and assume that the mean values of T and § are respectively proportional to the macroscopic octahedral shear stress Toct’ and octahedral shear strain rate Yoct.’ For un1ax1al loading, _ fz’ . _ . Toct. — -3'0' and Yoct. - J5 E where o and é are the applied axial stress and the axial rate of deformation respectively. Equation (3-2) can be rewritten in terms of these measurable'quantities, o and E, as 47 e = /2 %§ exp (%3) exp (- %%) sin h (%;w%%% 0) (3-3) Dorn and his coworkers (1954) investigated the variation of the structure parameter, Sp, with strain during creep at high temperature. They represented the creep rate of polycrystalline aluminum as '- -éfl- _ e — Sp exp ( RT) Sln h BO (3 4) If equation (3-4) is equivalent to equation (3-3), then S = /2 31 exp (%§) p h and B=Qfl 13 2kT For constant stress and temperature it seems reasonable to write either equation (3-3) or equation (3-4) as «I ll 2 SKZ' (3-5) where K2' is constant for a given stress and temperature, and S is defined as a structure term. This structure term may include such factors as softening of individual grains due to accumulation of strain along their basal planes (Steinemann, 1954), grain boundary creep, change 48 of grain size during creep, and other factors associated with ice deformation. nt Substituting él’ which is equal to Kl'e- , as proposed by Goughnour and Andersland (1968), and 62 from equation (3-5) into the time law equation (3-1), one obtains ' _ . -nt , _ e — Kl e + K2 S (3 6) Glen (1955) has discussed a similar argument (see page 33 ) as described by Figure 3-2. Steinemann (1954) also stated that if there were no recrystallization, then even after a slight macroscopic shear, some favorably oriented grains would be sheared by a sufficient amount to be soft, so creating a heterogeneous aggregate. The aggregate itself always constitutes an impediment to absolutely free glide of the grains and so clusters of soft grains would be the next stage. An effect of this sort may account for the observed accelerating stage (tertiary creep). On introducing recrystallization, an aggregate can remain hard if the recrystallization takes place in sufficient time to prevent the large glide, necessary for softening. If this recrystallization were absent a marked transition followed by a linear part would be expected. 49 If one assumes that softening of crystals within the polycrystalline mass is the dominant factor controlling the structure term, and S is defined to be the decimal per cent of softened crystals within the polycrystalline mass, then 01831. The rate of softening can be expressed as ' (S - So) (3-7) —. = ' . - - K 82(1 8) K4 where K3' and K4' are both constants for any given stress and temperature. So is constant and equals to S initially at time zero. The first term of equation (3-7), K3' 62(1-8), represents the assumed softening behavior, and the term K4' (S-SO) represents the effects of recrystallization. Substituting equation (3-5) into equation (3-7), one obtains 2 _d_§ _ _ g I l I _ I ' .- dt _ K2 K3 5 + (K2 K3 K4 )8 + K4 So (3 8) or rearranging a? 1 2 3 50 where Cl', C2' and C3' are defined as Cl — K2 K3 (3 10) C2 K2 K3 K4 (3 11) I _ I .— and C3 — K4 80' (3 12) K1' and K2' in equation (3-6) are defined as positive values. By definition K3' and K4' in equation (3-7) are also positive. Thus Cl' will be negative, C2' can be positive or negative, and C3' must be positive. After separating variables in equation (3-9), one obtains dS I I 2 S + C3 2 = dt. (3-13)‘ c 's + c l m = /(C2')2 - 4C1'C3' must always be real, therefore, integration of equation (3-13) gives 2C '8 + C ' + m i In 1 2 = t + c' (3-14) I I 2Cl S + C2 + m where C' is the integration constant. Equation (3-14) can be written as I I _ _ I _ 2Cl 81+ C2 + m = e mt mc = -ce mt (3_15) I I _ 2C1 S + C2 m 51 _ I where -C = e mC . The choice of signs was made so that empirical values of C and m will be positive. It is apparent that C' must be complex. Solving for S in equation (3-15), leads to I .. 2C ' . -mt 1 2Cl (1 + Ce ). As time approaches infinity, 8 approaches a steady-state or equilibrium value of C2' + m S = - -———-—-— (3-17) Now, substituting equation (3-16) into equation (3-6), one obtains I _ ' 1 2 2C ' . -mt 1 2C1 (1 + Ce ) or C ' -mt ° - t 2 n1 1 - Ce 8 = K 'e n - K . - K ' -——T ( _ ) 1 2 2Cl 2 2Cl 1 + Ce mt . Redefining the constants of the above equation, -nt 1 - ce'mt ( t) + K + K _m <3-18) l + Ce 1e 2 3 52 where _ I - Kl _ K1 , , (3 19) C I K2 =-K2' 35%7, and - (3-20) K =-K ' “‘ (3-21) 3 2 EEIT' By integrating equation (3-18), one obtains K 2K 1 e nt + K t + K t + -53 In (1 + Ce 2 3 s = - + C 7? s where C5 is the integration constant. And again redefining constants‘as - _£ _ C1 — 11 , (3 22) C2 = K2 + K3, and (3-23) 2K3 C3 = '—m-, (3-24) one finds _ _ -nt -mt _ e - cle + Czt + C3In (1 + Ce ) + cs. (3 25) Equation (3-25) is the final creep equation which is developed by the concepts described earlier. A series of tests were performed to test the applicability of equation (3-25). All parameters appearing in equation (3-25) can 53 be determined for any particular test by the computer least squares best fit method. Therefore, it is presumed that all parameters are known values. If at the beginning of the test, the strain and time both start from zero, equation (3-25) becomes -nt mt e = (—c e -+ci) + c tnt[C3In.(1rtCe- ) - C3In (1 + C)] 1 2 where ' (3-25a) c - C3In (1 + C) = c 1 5' If one takes the derivative of equation (3-25), the creep-rate will be given by é = P e"nt + P + P e-mt (3-26) 1 2 3 l + Ce—mt where P1 = Cln, P2 = C2, and P3 = -C3Cm. Equation (3-26) has the same meaning as equation (3-18) except all of the parameters appearing in equation (3-26) are known. Therefore, one can calculate the creep-rate at any time during creep. 54 From equations (3-24), (3-21) and (3-10), one finds K I Since C3 is known, K ' is known also. From equations 3 (3-23), (3-20), (3-24), (3-10) and (3-27) one obtains I I C=K+K=-K'C2+..—.—C3m.-=CzC3+——C3}m 2 2 3 2 2Cl' 2 2 2 ' This leads to .2C2 C2' = IT— - m. (3-28) 3 The value of C ' can be obtained from known values of C 2 2. C3 and m. In order to evaluate the constants, K2' and K4', the experimental procedure illustrated in Figure 3-3 is necessary. The sample is loaded to some stress which is maintained until time t1. The sample is then unloaded and allowed to "rest" until time t2, at which time the original stress is reapplied and creep continues. The first loading will be referred to as Phase I and the second loading as Phase II. Phase I is equivalent to a conventional creep test, and the parameters of equation (3-25) can be fit to Phase I test data, in the usual manner. Hypothetically equation (3-25) is valid for both 55 Unload Reload I I Phase I Phase II Strain ———" .‘_J&EEL2£M;_;J I Figure 3-3.--Schematic Diagram of Two Phase Loading. Phase I and Phase II. Consequently the parameters of equation (3-25) can also be fit to Phase II test data but these parameters will represent different initial condi- tions (80 value). At the beginning of loading in the first phase (t = 0), equation (3-16) can be written _ _ _ 2 _ m (l - C) _ S “ So ‘ 25:7 2c1' (I*+ C)’ (3 29) Multiplying equation (3-29) by KZ', and from equation (3'27) I 56 . __2_ m(l-C)___g_ _ K2 So “ 2 C3 + C3 E‘TI‘I'ET" 2 C3 + B (3 30) where B = C m'(1 - C) 3 2 (l + C)’ Substituting equation (3-28) into equation (3-30), one obtains (3-31) where KZ'So can be computed from known Phase I values of C C3, C and m. 2' During the rest period, while recrystallization is occurring, equation (3-7) should be written as = - K 010-- ('1’ U) 4' (S - So). (3-32) Integration of equation (3-32) gives In (so - S) = -K4't - c" or so _ s = e-K4't-C" = _ C'"e-K4't where C"' = -e-C" is constant. It follows that 57 -K4't S = C"'e + So. (3-33) Referring to Figure 3-3, when t equals t1, S equals S Then - K 't III = _ 4 1 Substituting back into equation (3-33), one obtains -K4'(t-t1) S = (Sl - So) e + 50' (3-34) When the rest time reaches t2, and S = $2, equation (3-34), multiplied by K2', becomes . , - 4'(:2-:1)I , — ' - - where t2 - tl is the rest time. K2' was originally assumed to be stress dependent, but since the stress is the same during the first and second phases, K2' is the same during the first and second loading phases. Now, from equation (3-16), one finds the expression of S in Phase I at t = t 1 1 58 . K ' Since from equation (3-27), C3 = - CETV multiplying by ' 1 K2' yields -mt C ' I , .V ..... 1 K2'51 = '2' C3 + C3 m (l ce-mE) 2 (1 + Ce 1) or it reduces to C2' I - _ K2 Sl — -§—'C3 + D (3 36) where -mtl _ m (l - Ce ) 2 (1 + Ce 1) If one substitutes equation (3-30) and (3-36) into equation (3-35), the result is -K '(t -t ) Kz's2 = (D-B)e 4 2 1 + -§ c ' + B. (3-37) Equation (3-37) expresses a relationship between Phase I and Phase II. If one considers time starting from zero in the second phase, it is possible to obtain best fit parameters of equation (3-25) for this data. Then for time zero for the second phase, equation (3-16) yields 59 K2 S2 2 C3 + C3 2 1+C) 2 C3 + B2 ‘3 38) _ In (l-C) Note that all parameters in equation (3-38) are result of Phase II test data. Equation (3-38) is equivalent to equation (3-37), thus C' -K'(t-t) C' 2 _ _ 4 2 1 2 —-2 C3|II+B2—(DB)e +——2 C3II+B (3-29) where the subscripts I and II indicate the Phase I and Phase II parameters. B2, D and B are defined as pre- viously. Since all parameters except K4' are known in equation (3-29), one can solve for K4', which gives C ' C ' 2 2 —-—cl --—c|+B-B K4' = t it In 2 3 II 2 3 I 2 (3_40) 1 2 D - B As long as K4' is determined, Kz' can be obtained from equations (3-11), (3-27) and (3-28), which gives I_. - .— K2 — K4' C3 + 2C2 C3m. (3 41) 60 K2' expresses the stress dependence of equation (3-5). More details of K "will be discussed with the experimental 2 results. Once K2' is obtained, Cl' can be found from equations (3-10) and (3-27). Thereafter the value of the structure term S can be found from equation (3-16) for any desired time. One can also obtain the constant So from equation (3-12) in which C ' can be found from 3 — 2- m - /(c2') 4 cl'c3' It is realized that So will be very small compared to S. As noted previously, when time approaches infinity, S approaches some equilibrium value. Equation (3-5) is then at the final creep rate stage, and equivalent to C2 in the equation of ‘ ‘ / This should be equivalent to equation (3-17). Extension to Multiaxial Creep In most analyses in Continuum Mechanics, two assumptions about the nature of the material are made, namely, that it is homogeneous, and isotropic. It should be clearly understood that there are actually three com- pletely independent assumptions. A material is continuous 61 if it completely fills the spaces that it occupies, leaving no pores or empty spaces. A homogeneous material has identical properties at all points, and a material is called isotropic with respect to certain properties if these properties are the same in all directions. The determination of the continuum concept of matter for ice has to be investigated. A thin-walled, hollow, cylindrical ice sample was considered for this study. This model can be applied to three dimensional stress states. At any point in the material, assumed to be homo- geneous and weightless, let Oij be the components of stress, and Ui be the components of flow. The equations of slow motion of an element are then 80. fi=O(i.j=l,2.3) where the summation convention is used for repeated suffixes (Nye, 1953). The rate of strain tensor at any point has components , 1 3U]: 3U. Eij = '2' (3;; + ‘43in- (3'43) For incompressible material, i.e., 81 + 62 + £3 = 0, and isotrOpic throughout the flow, the principal axes of éij must be parallel to those of Oij and the 62 components of the strain rate proportional to the components of the stress deviator, thus 8.. = 10!. (3-44) 13 13 where Gij' the stress deviator, is given by 0 =0 --1-3..c (3-45) ij 13 3 13 kk and A is a positive scalar factor of proportionality that depends on position and time. The term %.aij Okk is called the spherical stress tensor, and can be considered to act as a hydrostatic stress. In plastic flow considera- tions one can consider the stress system obtained by sub- tracting the spherical state of stress from the actual state, rather than working with the actual state of stress. There is furthermore evidence that the hydrostatic pressure does not affect the flow rate of ice (Rigsby, 1958). In three-dimensional stress states, the shear stresses and shear strains on the octahedral plane give a good representation of the shear distortion in the specimen. If experimental stresses and strains are taken under the principal directions, they can be converted to octahedral stresses and strains. The theoretical creep equation as derived in equation (3-25) must be modified then to be applied to the thin-walled, hollow, cylindrical samples. 63 The octahedral shear stress is given by _ 1 _ 2 _ 2 _ 2 _ Toct. - 3 H01 02) + (02 03) + (a3 cl) (3 46) where 01, 02 and 03 are three principal stresses, and octahedral shear strain is given by _ 2 _ 2 _ 2 _ 2 _ Yoct. — 3 /(e1 :2) + (62 63) + (e3 61) (3 47) where 81, £2 and 83 are three principal strains. In uniaxial loading, equation (3-46) becomes _ J? Toct. - IT 0' (3 48) and equation (3-47) becomes 7 = /2 e I (3-49) oct. where c and 5 denote the applied axial stress and the axial strain respectively. In order to convert equation (3-25) from the principal sense into the octahedral sense one must multiply by a factor of /2 and convert the stress dependent param- . 3 . eters from 0 into /E T . The result 13 oct. 64 -noct t eoct. = -Cl oct. e + C2 oct. -moct.t (1 + Ce ) + c5 oct ' (3-50) One can determine the best fit parameters of equation (3-50) for experimental data by using the same technique as described previously. Theoretically, these parameters determined from the results of tests on thin-walled, hollow, cylindrical samples should be identical to the results of solid, cylindrical samples. The comparison between these two results will be given later with experimental results. Plastic Strain Computation Solid, Cylindrical Sample When a solid is subjected to a static, axial force, the atomic lattice will adjust itself to oppose the applied force and maintain equilibrium. On a macroscopic scale the atomic adjustment is observed as deformation when the lattice remains continuous and as fracture when the atoms are pulled apart. Deformation involves displace- ments which when sufficiently large can be measured macroscopically. The magnitude of these displacements depends on the dimensions of the element deformed. Dis- placements cannot be compared directly unless referred to unit elements and thus converted into a dimensionless quantity called "strain." Under a simple tensile or 65 compressive force the strain can be directly computed from displacement, or change in length, in the direction of applied force. Often the strain is computed as a normal strain E, which refers the change in length, dk, to an original length, 20, and is given by _ 2 . =I Tet—l =37 (3-51) The displacement also can be converted into a true strain, a, which refers to the change in length to the instantaneous length and is given by z _ From equations (3-51) and (3-52), one obtains e = III (1 + E). (3-53) For compression, equation (3-53) becomes m ll - In (1 - E) or - In (1 - A31). (3-54) 00 ll 66 If one only measures the plastic deformation of the speci- men, subtracting elastic deformation and system error during the course of the experiment, equation (3-54) can be written as . A9; ._. C ILo e = -In [1 - ] - G (3-55) where C is elastic deformation, and G is the system error which is calibrated from the test equipment. Equation (3-55) was used to compute the plastic strain for all uniaxial compressive creep tests in this study. Thin-Walled,iHollow, Cylindrical Samples When pressure is applied internally to a thin- walled, hollow, cylindrical sample, the experiment is con- sidered to be a tensile test since the specimen is radially expanded. The conventional strain can be written as r.- = —fi—2 = g2 (3.56) o o where D0 = initial inner diameter of the hollow cylinder, and D = instantaneous inner diameter of the hollow cylinder. 67 The initial volume and current volume can be expressed as nDoz v0 = 4 hO ‘ (3-57) 2 _ flD _ V _.Tho (3 58) Where ho is the height of specimen. Substituting equations (3-57) and (3-58) into equation (3-56) yields _ = (_y_)1/2 5 v 0 _ 1 (3-59) Applying equation (3-53), the true strain becomes a = %- In (3-60) oHsO mmmuuulénm 053m . Cw: mafia OOH om cm on. om om 0v om ON OH q u _ q . u q q — _ A land ommv muszz mm.~ Uomqul ll BO em. 5 4mm OH NH 3” ma ma % utexns IPIXV 112 HN ON .mv mHmEmm .cflmnum Hoaxd momHm> mummIsflmuum HmwdeI.NIm whomflm ma ma ha 0H mH va a seesaw amass ma NH HH OH a . lane ommo q q d N :22: He.H o Uomv.vl 8 me wadsmm - d W s s 8388 urezns IPIXV m.H _ m.H ..Ioaxo.~ Axial Strain % Axial Strain % 16 14 12 10 30 113 (a) Showing Three Loading Cycles Rest 7.344 x 104 5 Rest 7.764 x 104 s l 1 1 1 l 1 l l J l I l L 60 90 Time min. (b) Loading Cycles Superimposed 3rd Loading S lst Loading ample 6A T = -4.45°C o‘ = 1.715 mun'z 2nd Loading 1 J l 1 1 I 1 1 J I l l l l 120 150 180 210 240 270 '300 330 360 '390 '420 450 (249 psi) 10 20 3O 4O 50 60 70 80 90 100 110 120 130 140 Time min. Figure 6-3.--Creep Curve for Sample 6A. 114 land ooav .monEmm m mmHHmm Hmuw>mm How m>uoo mmmuonu.va wusmHm OH x NH HH OH .sez mass N .22: mmo.o n o Uom¢ .fil " .H. ma maesmm Ase OH NH VH 0H mH om % urezns Isrxv .pmssHusooII.va wuomHm .sez wees OH x 0N vN NN 0N mH mH vH NH OH m o v II 0 land omal muezz omo.H 115 .H. Uomv.vl mm mamssm Any OH NH vH 0H mH ON % Utezns IPIXV 116 668.388....78 053s .GHS 08:. 3x2. mo 8 mm om ms 9. mm on mm on 3 S _ _ _ e a _ . e a . _ H . 000.. o o o o 0 0 land come mIszz mom.e n p oome.su n a 6 mm seesaw on O OH NH VH 0H mH ON % “rains TPTXV .OmocHusooII.¢Io musmHm 117 .sa: mass can omm oo~ omH ooe om _ H _ _ _ . niIm o o‘w‘o O o o o o 0 o O 0 O O O O o 0 land ommo szz He.a u o o o NI Uomv.fil fl 8 mm was me nmamssm any 0 O «H OH OH ON NN. % uthqs Isrxv 118 683:8). ens 083m .cHS OEHB ONH OOH OmH OVH OMH ONH OHH OOH om Om on 00 Om 0v Om . _ _ H _ _ H _ H H . _ H _ q me seesaw .leme cove Nuszz owe.~ mm «Assam .lene 0mm. .32: omm.~ ms seesaw .ienm come «.22: omo.~ mo seesaw .xenm oomo Nuszz ooo.~ Uomfiofll H mm psa..mm .me .mo nweessm Amy I B D O b D O mh.v mm OH NH VH OH OH ON % ntvzns I9? Residuals % Axial Strain % 22 20 18 16 14 12 10 119 10 20 l 30 I 40 Sample BB T - -4.45°C o = 2.38 MNm-2 (350 psi) I l l l l 50 60 70 80 90 Time min. Figure 6-5.--Creep Curve for Sample 8B Showing Correlation with Equation (3-25). 120 .mmmuum mo QOHuosom 0 mm HO HO mmsHm> uHm umwmul.oIo muomHm szz museum H602 m.N O.N m.H O.H m . u —I - H d J) q n J )1 d - d d + q — 1 4 d d — - I1 0 O G e w Hmmo .1. 09395 O O G _ .mmmuum HMde msmum> Ammlmv :oHumsom Eonm 0 mo mmsHm> uHm ummmII.hIo whomHm szz mmmflm H302 m.~ o.~ m4 o; m. o 121 O G . u mmmuwtsn mu m NO ioOH IONH IOVH 1OOH 122 - O 10-2 r C) )- 10'3 — C. H- l m)— a 5 4 8 1 ' n = 8.28 x 10- O ' 5 s 104.. r- - 1 1 l LIlllllllIIlllllllliil 10 0.5 - 1.0 1.5 2.0 2.5 3.0 Axial Stress MNm"2 Figure 6-8.--Best Fit Values of n from Equation (3-25) Versus Axial Stress. 123 10 IIIII—I 10 10 r TT—III' I ilIiJIIIIllililIilll 0.5 1.0 1.5 2.0 2.5 3.0 Axial Stress MNm- Figure 6-9.-Best Fit Values of m frun Equation (3-25) versus Axial Stress. 124 10.2 C I. 10"3 . I. H I- I m I- N r- N U 10'4 - P 10-5 1 l l l lllllLlllIlLLJlIllLII 0.5 1.0 1.5 2.0 2.5 3.0 Axial Stress MNm-2 Figure 6-10.--Best Fit Values of C2 from Equation (3-25) Versus Axial Stress. 125 .nmmuum amass momuo> Ammumv :oHumoom scum m0 m0 moon> uHm ummmII.HHIm musmHh NLEZZ mmwuum HMme o.m m.N O.N m.H O.H m.o H H _ _ s mm.mu s em.m u no my m mO.N u U Axial Strain % 126 40 b 30 — 20 7 10 .. I 0 J l 1 0 1 2 3 ‘Axial Stress MNm"2 Figure 6-12.--Axial Strain Necessary for Softening Term to Reach Stated Per Cent of its Original Value as a Function of Stress. mcfimmmuocH haumsnn¢ mo OHXm m e m . m .cflz mafia o moans m uommmm mcwzonm moa mamfimm Mom mm>udu mwmuo||.malo musmwm 127 1 m sea x Hm.m u w Hanan Aflmm cemv muszz no.~ u o mmmnm mcwwmoq vacuum any a 1 Aflmm cows wuazz mm.H u o mmmnm mcfiomoq umuflm Ame OH % Utezus I9? .omoq.mcwmmmuomo maumaund Mo uommmm mcfl3onm mHH mHmEmm How mm>uso mmmuoul.valm musmfim 128 NOH x ow mm on mm ON m.” O.H m .cez mafia m2 o3 m3 2: ms om mu 0 _ A q . _ _ _ q fil w .wmm made .222 mmo.o u o N. mmmsm oceanoq ccoomm Any q a a Aflmm come NIEZZ ho.N u .o mmmnm maflomoq pmuflm Amy ON mm % urvxns IEva 129 .mcwcmoq wmmnm,ccoumm ca mma can med .mma .mNH mmamfimm How mm>uco mmmnunn.maum musmwm .5": mafia. omN CON cm...” 00.... om o A e — - , _ q 0“: 4V 0 D Ad 0 d o J as o J 0 l Acme ommv muazz H>.H u o I UomVoQI an H. q o unwvmoq mmmsm wcoomm d 0 4 o 1 q o mma J OH NH 3.. 0H ma ON NN % uxezns IPTXV 130 10'1 _ L- -2 10 b , _ -4 . 0 : K2 — 3.68 x 10 Sin h 6:353- 01 w i -N __ M 10'3 _ y- 10-4 L l 1 njlnnllllnilijjilninll 0.5 1.0 1.5 2 2.0 2.5 3.0 Axial Stress MNm— Figure 6.16.--Values of K2' Versus Axial Stress. 131 l E L. m #- U) 1o'i P- 10- L 1 l 1 [111L141L4J1111111111 0.5 1.0 1.5 2.0 2.5 3.0 Axial Stress MNm—2 Figure 6-17.--Plot of Equilibrium Values of S Versus Axial Stress. 10"2 f r L 10"3 t m I w I- O .. U) -~ _ M B” F J . U) 'N M 10" r ; p. 10"5 : 0.5 132 Stress. 1.0 1.5 2.0 Axial Stress MNm"2 Figure 6-18.--Computed K2'Ss, Cz' and K 2 I So Versus 133 .mma mamfimm 00H mafiaamummHUMHom How m>uso mmmHUII.mHIo shaman musom mafia omm omv ovw 00¢ 8m omm omN ovN DON om.” ONH om 0v 0 J _ . q _ _ a q _ d _ d d N.o m.o v.0 00m K . a- macs ma N lama mmc «uszz mmma.o Uomvofil u u 0w m.o E-‘b mod mamsmm % urexns IPIXV Axial Strain Rate 5 10 I YI‘TIIl r 10 vrxl‘ I 10 ,. 10 #11 TT‘Vf'l 10 I fir!!! T 134 = -2.06°C Sample 2H Result Sample 168 Result I 1 J 1 0.1 ‘ 1.0 10 100. Axial Stress Kgf/cm2 Figure 6-20.--Axial Strain-Rate Versus Axial Stress (After Mellor and Testa, 1969). 135 CA N mm .mH mamfimm How 0>HSU mmmHUII.HNIm musmflh .cfiz mafia om mv 0v mm on mm ON mu OH m o 00m. w HI lame moNV m MI. _ . q _ . _ . _ . 7 _ OH x mmo.H u w .52: Hee.H use come.v- u a ma mflmsmm OH NH 2. 0H ma % urexns terxv 136 OHN mH .Aumme mmmuo mHstmBV mm mHmEmm sou m>uso mmmuunl.mmlm wusmHm .cHz mEHB ¢H mH NH HH 0H m m n o m v m N H o 1 d _ . - H _ H _ — 4 0mm x . u u HI N...0H mH H . lame «as «:32: oam.o u no Uomv.¢l u .H. L umma mwmuo wHHmsma mm mamsmm 1 L % urexns IPTXV 137 .5 398m now 9.50 m8uul.mmlm musmHm .52 22. «S x Z 3 NH S 3 m n o m e m m H o — u — q - d fi — d A u d O)<“l°..0.c q 4 O O O 1 O a. q 0 O 0 q o C 0 00m OH x Nv.H Imw a 0 HI 0| I o d a c Q coHumEHomwa Hmuwug O O o Ream code «-52: mmm.o u mo . .. m a can 0 ca x ~e.m "Hm o lame vac «-52: memo o 1 Ho . o isms «may muszz «mm.H u o Uomfiofil “ E 0 mm «38mm 0 “83.9.50me Hausa 0H NH. vH 0H. mH % uIexns IEIXV Octahedral Shear Strain % 4O 35 3O l 2 3 4 S 6 '7 8 9 10 138 Sample 3H T = -4.4S°C T = 0.501 MNm-2 (72.7 psi) .oct. _4 _1 e = 7.3 x 10 % sec oct. L l 1 l l l l l l l l L l L ’ 2 11 12 13 14 x 10 Time Min. Figure 6-24.--Octahedra1 Shear Strain Versus Time for Sample 3H. 139 10"1 f -2 10 L h -3 4.85 _ oct. - 3.15 x 10 (Toot.) , P m .. J o J) . 10”3 - ' Sample 3H Result -4 10 l 11 #1111 1 1 JlLlllJ 0.1 , 1 10 Octahedral Shear Stress MNm.-2 Figure 6-25.--Plot of “oat Versus Octahedral Shear Stress for Samples 1B through 9B Compared with Sample 3H. 140 10-2 r r _ -3 4.05 OCt. — 1.22 x 10 (Toct.) 10'3 - F m b 43 U 0 I- 5 Sample 3H Result 10"4 - P r 10-5 1 lJJLlLl 1 1 11111LJ 0.1 1.0 10 Octahedral Shear Stress MNm—2 Figure 2-26.--Plot of moct Versus Octahedral Shear Stress for Samples 1B through 98 Compared with Sample 3H. 10 10 C2 oct. % S 10 TTUIj rrlli1 Figure 6-27.-Plot of C 141 _ -3 4.22 - 6.36 x 10 (root) C2 oct. 3H Result 11L141l 1 1 141111] 1.0 p 10 Octahedral Shear Stress MNm.-2 2 col: Versus Octahedral Shear Stress for Samples 1B through 9B Cmpared with Sanple 3H. 142' 1 k .mm mHmEmm How mummlsHmuum can HoumH>ma mmwnum mo muouom> mo coHuwucmmmumwm UHpmfimnom11.m~1m musmHm EXT m . o H H H IIIIIIIIIIIIIIIII Inllllllh H H H H IIIIIIII MN : u __ _ [8| _ _ _ _ _ _ _ _ _ [Dbl __ _ _ m _ __ w IIIIII _L$MJ. I \ _ \J _ _ \\ _ _ .\ _ _ _ x... -u- .ewmmvé- " _ _ _ _ _ _ _ _ _ . __ _ _ YLLovu _ __ _ _ . 09. __ _ _ n u %v 60 . _ . _ 31.8.. _ r .w . __ n u . _ +A..r my _ _ _ _ _ a r... __ _ _ NITTONI _ m __ _ _ . _ 9 _. . _ _ _ é __ . \r---- _-.2.. o _ _ _ \ _ f 1.3% in x... .... _. X m .1 . Ell- .l' —III a I d Ill-'Ilqllllq1111l' V is as om am a» .8. an. 8 on 8 2 )w s . \ 0’ a + \ Q? A? % same \ com 9 AH: my .mo 143 OH x mm.H m.mm OO.m OH x mO.H OH x hm.w wem.O oov Owh.~ mm m1 m1 m1 mIOH x mm.H m.mh ow.v mIOH x mu.h mIOH x -.O «Hm.o omm OOMwN mm ¢IOH x hm.h O.mb va.v mIOH x O~.¢ mIOH x mm.m vmm.o com OOO.~ . ms «IOH x hm.m h.mm OO.> mIOH x Oh.¢ m1OH x h¢.v oev.o oom OOO.~ mm «IOH x N>.¢ m.vm mm.m m1OH x mm.H mIOH x mo.~ mmm.o omm OHn.H mm «IOH x NH.m O.mm mm.~ mIOH x mh.H mIOH x mO.m ovm.o omm OHh.H me enoH x Om.~ N.S¢H vo.~ eIOH x -.h vIOH x Oh.m MN0.0 OON mom.H mm mIOH x em.w ¢.mo mh.~ vIOH x w~.m vIOH x HH.H omm.o omH omo.H mm mIOH x H~.H v.mv Hm.~ mIOH x NH.v mIOH x OH.H OO0.0 OOH Omo.O mH Hum M Hum m Hum Hm Hmm WIEZZ .oz 2 o o o a 0 39:8 mmmuum HMHX¢ .Uomvofil " B .mumme m mmHme on uHm ommm How owumsHm>m Ammlmv cOHumswm mo mumqumumm11.H mqm u can Ho cuss .mumwa m moHHmm ou uHm umwm MOM omumsHm>m Ammlmv QOHumswm mo muwuofimumm11.~ mamas 145 v m IOmemm.m u . M mommm>¢ H1 v vIOH x Nmm.m vIOH x om.o m.mm mm.m MIOH x mm.H MIOH x mo.m mmm.o oomh MH.v mm.v comb mm vIOH x whm.m vnOH x hm.w O.mo mv.H mIOH x Nh.H MIOH x vo.¢ mov.o oovm mH.v vm.v comb mvH vIOH x vhv.m vIOH x mm.h m.OH vo.~ MIOH x om.H mIOH x vm.m mom.o ooom Om.g om.v OVHS mm vIOH x Omm.m «10H x mm.w mm.m mv.v MIOH x mm.m NIOH x ov.H mmm.o oowH mH.¢ mm.v comb mN m m a ma w w w m Hlv H1 E 0 mo H1 NU Hm Ho m cwmuum chuum made um. . M wade Omusmeoo Hmouom . . cmusmfiou mmmam OQHUMOA ocoomm ummm mmmnm OCHGMOA umHHh v .Asmm ommv «-52: He.a mo mmmuum paw oomv.v1 um mumme HH¢ .. M ucmumcoo coHHMNHHHmumhnomm mo wsHm> wcHEHmswa ou mamas mo muHSmmmI1.m mqm, 146 ¢IOH x O.m mIOH mmm.v MIOH x mum.H mIOH x hmm.mu OHH.O mmmm.o NuoH x mo.N vh.~ mm vIOH x m.m vIOH mHN.H MIOH x mmm.H MIOH x SOH.NI mHm.O thw.o NIOH x HO.H mm.~ mm vIOH x 0.0 mIOH Nvm.b vIOH x wo>.m MIOH x mom.H1 OON.O mHm0.0 MIOH x mm.m oo.~ mu «10H x m.m mIOH mmw.H vIOH x mmmro vIOH x th.m1 ONH.O wmmm.o mnoH x Hm.h OO.N mm «10H x m.m mIOH mnm.m vIOH x mnm.m vIOH x mm~.m1 mmm.o omnm.o maoH x mm.m Hh.H mm vIOH x ¢.~ mIOH mom.m «IOH x mvm.o m10H x nNO.H1 mvm.o nmmm.o mIOH x OO.m Hh.H mo «IOH x mv.H mIOH me.m vIOH x mHN.v v10H x ovm.hu omv.o vmmv.o mIOH x vm.H mom.H mm mIOH x b.m wuoH mom.m mIOH x vov.m «IOH x mo~.v1 mmm.o nHmH.o mIOH x mH.H mO.H mm OIOH x m.h h10H th.H mIOH x NNh.H @10H x www.mn 0mm.o omeo.o vIOH x mm.o omm.o mH AHmumO N Ammnmv AOHIMO Anmlmv N m AHvumv 52: .oz coHumsvm A E: . UYIII u :oHumsvm coHumsUm coHumswm . M\ U u codumswm N1 N , . mmwuum onEMm mm m m 1m w m w 07m Hm Hm I T 7 Have W. Vm - U . U a O .MVH mm .Nx .momme 0 muss o Ho 0 N m . N . H . v N .11 um m m w m. M can . U . U . U . M .mm .. M NO mwsHm>Il.v mqmda 147 TABLE 5.--Value of S as a Function of Stress and Time. Axial Sample Stress S - S S No. MNm-2 T=3000 sec. T=6000 sec. T=9000 sec. 1B 0.686 0.00809 0.00812 0.00814 2B 1.03 0.0319 0.0326 0.0334 3B 1.365 0.0961 0.10389 0.1211 43 1.71 0.1029 0.1831 0.3568 SE 1.71 0.1066 0.1939 0.3583 6B 2.06 0.1381 0.3645 0.5561 73 2.06 0.1581 0.3930 0.6019 SB 2.38 0.4710 0.7685 0.7740 93 2.74 0.7177 0.8291 0.8296 148 momuo HMde mauve «IOH x N.N m.mm mm.“ m..OH x OO.H vIOH x O.m Hmm.o va.H m moHuom ¢IOH x Onm.H hm.Om bm¢.> MIOH x omO.H. MIOH x VH.H bmv.o va.H mmwuo Hme< mH w m w m w 522 . H1 H1 H1 N1 9 oz E 0 m0 NU : Ho mmouum m as umma mHmEmm .mH mflmsmm op yum ummm you emumaam>m Ammumv coaumsem «0 mumumfioummll.o mamas 149 OH x O¢.h «.mm mh.m OH x ¢¢.m IOH N MOH.H NN0.0 Hom.O mmwuum m1 «1 w muMHumoummm How mummu m meumm «IOH x m~.H me.om me.m euoa x om.h «nos x ~m.H oe.H Hom.o mm m o 00 m E H- u. o a H- m Hum w ~._zz moz . . . . . mmmuum ma saw #008 “.00 MO “.00 1H0 U00: “00 Ho Hmmgm goo .mgHSmmm m mmHHmm suH3 oouwmfiou mm mHmEmm op uHm ummm MOM omumus>m mnmumamumm Hmuomsmu0011.h mqmda CHAPTER VII CONCLUSIONS The conclusions summarized below are based on the observed results of experiments conducted on samples of polycrystalline ice at a temperature of -4.45°C and the theoretical analysis presented in this study. 1. Equation (3-25) was derived by a hypothesis which describes the creep behavior of polycrystalline ice in terms of the mobilization of grain boundary effects, softening of individual crystals due to accumulated strain on their basal planes, and recrystallization. It appears to account for this behavior through primary, secondary and tertiary stages of creep for the range of stresses and temperature studied. The equation predicts a creep-rate that approaches some constant value when grain boundary effects are fully developed and as the quantity S approaches some equilibrium value, where S is defined as the decimal per cent of soft crystals in the polycrystalline mass. Experimental results from uniaxial creep tests and multi- axial creep tests for high stress levels appear to fit this form of equation for strains up to at least 25 per cent. 150 151 2. It is observed from experimental results that approximately 10 per cent of total strain is necessary to reach the tertiary steady state creep. In other words, approximately 10 per cent axial strain must occur in poly- crystalline ice before all crystals have opportunity to soften. This is in agreement with results obtained by Steinemann (1954). 3. A value of about 20 Kcal/mole for activation energy was found for sample tested at a temperature -4.45°C. This value agrees favorably with results by previous experimenters. 4. The results from creep tests under low stress seem not to satisfy the theoretical creep equation (3-25). Since steady-state tertiary creep was not developed for these tests, application to the equation (3-25) is not valid. However, the secondary creep-rate from these tests agrees well with previous results by Mellor and Testa (1969). The lack of a tertiary creep stage may be due to the fact that these tests were not continued to a total strain of 10 per cent. 5. The validity of the assumption by Nye (1953) that for polycrystalline ice the principal strain rates are parallel and proportional to the corresponding princi- pal stress deviators was verified by the tests on thin- walled, hollow, cylindrical ice samples. The results were “A.“ - 152 satisfactory. Discrepancies are attributed to impurity of ice material (nonhomogeneous and anisotropic) or experimental errors. BIBLIOGRAPHY 153 BIBLIOGRAPHY Bridgman, P. W. "Water, in the Liquid and Five Solid Forms, Bromer, under Pressure." American Academ of Arts and Science, Proc. 47( , , pp. -558. D. J., and Kingery, W. D. "Flow of Polycrystalline Ice at Low Stresses and Small Strains." J. of Applied Physics, 39(3), 1968, pp. 1688-1961. Butkovich, T. R., and Landauer, J. K. "The Flow Law for Ice." Research Rpt. 56, U.S. Army Snow, Ice and Permafrost Research Establishment, August, 1959. Butkovich, T. R., and Landauer, J. K. "Creep of Ice at Low Stresses." Research Rpt. 72, U.S. Army Snow, Ice and Permafrost Research Establishment, August, 1960. Chalmers, B. "Micro-Plasticity in Crystals of Tin." Proc. Coble, Conrad, Royal Soc. of London, Ser. A. 156, 1936, pp. 427- 443. R. L. "A Model for Boundary Diffusion Controlled Creep in Polycrystalline Materials." J. of Applied Physics, 34, 1963, p. 1679. H. "Experimental Evaluation of Creep and Stress Rupture." Chapter 8, Mechanical Behavior of Materials at Elevated Temperatures, ed.‘By J. E. Dorn. McGraw-Hill Book Co., Inc., New York, 1961, p. 149. Cottrell, A. H., and Jaswon, M. A. "Distribution of Solute Dillon, Atoms Round a Slow Dislocation." Proc. Royal Soc. of London, Ser. A, 199, 1949, pp. I64-114. H. B., and Andersland, O. B. "Deformation Rates of Polycrystalline Ice." Internat. Conf. on Physics of Snow and Ice, The Institute of Low Temperature Science, Hokkaido Univ., Sapporo, Japan, August, 1967. 154 155 Dorn, J. E. "Some Fundamental Experiments on High Tempera- ture Creep." J. of Mechanics and Physics of Solids, 3, 1954, pp. 85-116. Dorsey, N. E. Properties of Ordinary Water-Substance. Reinhold Publishing Co., New York, 1940. Friedel, J. "Dislocation Interactions and Internal Strains in Internal Stresses and Fatigue in Metals." G. Rassweiler and W. L. Grube, eds. Elsevier Publishing Co., Amsterdam, 1959, pp. 220-262. Garofalo, F. Fundamentals_of Creep and Creep-Rupture in Metal. Macmillan Sefiés in Material Science, I965. Glasstone, 8.; Laidler, K. J.; and Eyring, H. The Theor of Rate Processes. McGraw-Hill Book Co., Inc., New York,II§41. Glen, J. W. "The Creep of Polycrystalline Ice." Proc. Royal Soc. of London, Ser. A, 228, 1955, pp. 519-538. Glen, J. W., and Perutz, M. F. "The Growth and Deformation of Ice Crystals." J. of Glaciology, 2(16), 1954, pp. 397-403. Gold, L. W. "Some Observations on the Dependence of Strain on Stress for Ice." Canadian J. of Physics, 36, 1958, pp. 1265-1275. Gold, L. W. "Deformation Mechanisms in Ice." Chapter 2, Ice and Snow, Properties, Processes and A lica- tions, ed: by W. D. Kingery. MIT Press, CafiSridge, Mass., 1963. Goughnour, R. R. "The Soil-Ice System and the Shear Strength of Frozen Soils." Ph.D. thesis, Michigan State Univ., East Lansing, Mich., 1967. Goughnour, R. R., and Andersland, O. B. "Mechanical Properties of a Sand-Ice System." J. of the Soil Mechanics and Foundation Div., ASCE, 9413M 4), Proc. Paper 6040, 1968, pp. 923-950. Griggs, D. T., and Coles, N. E. "Creep of Single Crystals of Ice." Research Rpt. ll, U.S. Army Snow, Ice and Permafrost Research Establishment, 1954. Halvorsen, L. K. "Determination of the Modulus of ' Elasticity of Artificial Snow-Ice in Flexure." Research Rpt. 31, U.S. Army Snow, Ice and Permafrost Research Establishment, February, 1959. 156 Hawkes, L. "Some Notes on the Structure and Flow of Ice." GeOl. Mag. ' 67’ 1930, pp. 111-1130 Herrin, M., and Jones, G. E. "The Behavior of Bituminous Material from the Viewpoint of the Absolute Rate Theory." Proc. Association of Asphalt Paving Technologist, 32, 1963, pp. 82-101. Herring, C. "Diffusional Viscosity of a Polycrystalline Solid." J. of Applied Physics, 21, 1950, pp. 437- 445. Higashi, A., and Sakai, N. "Movement of Small Angle Boundary of Ice Crystal." J. Faculty of Science, Hokkaido University, II, 1961} Itagaki, K. "Self-Diffusion in Ice Single Crystals." Research Rpt. 178, U.S. Army Material Command, CoRoRoEoI-lo' 19660 Jellinek, H. H. G. "Plastic Deformation of Thick-Walled Snow-Ice Cylinder under Hydrostatic Pressure." Research Rpt. 63, U.S. Army Snow, Ice and Perma- frost Research Establishment, 1960. Jellinek, H. H. G., and Brill, R. "Viscoelastic Properties of Ice." J. of Applied Physics, 27(10), 1956, pp. 1198-1209. Kamb, W. B. "The Glide Direction in Ice." J. of Glaciology, 3(30), 1961, pp. 1097-1106. Kanter, J. J. "The Problem of the Temperature Coefficient of Tensile Creep Rate." American Institute of Mining and Megallurgical Epg., Trans., 131, 1938, pp. 385-404. Kauzmann, W. "Flow of Solid Metals from the Standpoint of the Chemical Rate Theory." Trans. Am. Inst. of Mining and Metalurgical Engin., 143, 1941, pp. 57-83. Krausz, A. S. "The Creep of Ice in Bending." Canadian J. of Physics, 41(1), 1963), pp. 167-177. Leonards, G. A., and Andersland, O. B. "The Clay Water System and Shearing Resistance of Clays." Research Conference on the Shearing Strength of Cohesive Soils, Soil Mech. and Found. Div, ASCE, Univ. of Colorado, Boulder, Colo., June, 1960, pp. 793-817. 157 Levi, L.; DeAchaval, E. M.; and Suraski, E. "Experimental Study of Non-basal Dislocations in Ice Crystals." J. of Glaciology, 5, 1965, pp. 691-698. Lonsdale, K. "The Specific Gravity of Ice." J. of Glaciology, 1, 1950, p. 442. Main, J. F. "Note on Some EXperiments on the Viscosity of Ice." Proc. RoyppSoc. of London, 42, 1887, pp. 491-501. McConnel, J. C. "On the Plasticity of an Ice Crystal." Proc. Roy. Soc. of London, 49, 1891, pp. 323-342. McConnel, J. C., and Kidd, D. A. "On the Plasticity of Glacier and Other Ice." Proc. Roy: Soc. of London, 44, 1888, pp. 331-367. Mellor, M., and Testa, R. "Creep of Ice Under Low Stress." J. of Glaciology, 8(52), 1969, pp. 147-152. Muguruma, Jo, and Higashi, A- "Observation of Etch Channels on the (0001) Plane of Ice Crystals Produced by Nonbasal Glide." J. of Physics, Soc. Japan, 18, 1963, p. 1261. Nabarro, F. R. N. "Deformation of Crystals by the Motion of Single Irons." in Report of a Conference on the Strength of Solids, the Physical Society, London, 1948, pp. 75-90. . Nakaya, U. "Mechanical PrOperties of Single Crystals of Ice." Research Rpt. 28, U.S. Army Snow, Ice and Permafrost Research Establishment, October, 1958. Nakaya, U. "Viscoelastic Properties of Snow and Ice in the Greenland Ice Cap." Research Rpt. 46, U.S. Army Snow, Ice and Permafrost Research Establish- ment, May, 1959. Nye, J. F. "The Flow of Ice from Measurements in Glacier Tunnels, Laboratory Experiments, and the Jung- fraufirn Borehole Experiment." Proc. Roy. Soc. of London, Ser. A, 219, 1953, pp. 477H489. Pounder, E. R. The Physics of Ice. Pergamon Press, Oxford, 1965. 158 Rigsby, G. P. "Crystal Orientation in Glacier and in Experimentally Deformed Ice." J. of Glaciology, 3, 1960, pp. 589-596. Rigsby, G. P. "Effect of Hydrostatic Pressure on Shear Deformation of Single Ice Crystals." J. of Glaciology, 3, 1958, pp. 273-278. Shoumsky, P. A. "The Mechanism of Ice Straining and Its Recrystallization." IUGG, IASH, Chamonix Shumskii, P. A. Principles of Structural Glaciology. Dover Puincations, Inc., New York, 1964. Steinemann, S. "Results of Preliminary Experiments on the Plasticity of Ice Crystals." J. of Glaciology, 2, 1954, pp. 404-412. Van Vlack, L. H. Elements of Materials Science. ' Addison-Wesley Publishing Co., Inc., Reading, Mass., 1959. Weertman, J. "The Eshelby-Schoeck Viscous Dislocation Damping Mechanism Applied to the Steady-State Creep of Ice." Ice and Snow, Properties, Processes apd.App1ications, ed. by W. D. Kingery, MIT Press, Cambridge, Mass., 1963, pp. 28-33. APPENDIX DATA 159 160 SAMPLE 1A SAMPLE 2A T = -4.45°C T = -4.45°C Y1 = 0.902 gm/cc Yi = 0.905 gm/cc o = 350 psi . V..,o = 350 psi ....... Time Deflection Time Deflection (min.) (in.) (min.),, .(in-).. 0 0.0076 0 0.0076 1 0.0126 1 0.0122 3 0.0173 3 0.0182 5 0.0206 5 0.0224 10 0.0274 10 0.0337 15 0.0344 15 0.0412 20 0.0418 20 0.0523 25 0.0508 25 0.0660 30 0.0620 30 0.0836 35 0.0760 35 0.1054 40 0.0938 40 0.1326 45 0.1162 45 0.1638 50 0.1439 50 0.1984 55 0.1766 55 0.2369 60 0.2138 60 0.2782 65 0.2550 65 0.3228 70 0.2982 70 0.3715 75 0.3441 71 0.3822 Load Removed Load Removed SAMPLE 3A T = -4.45°C 71 = 0.904 gm/cc o = 350 psi Time Deflection (min.) (in.) 0 0.0080 1 0.0134 3 0.0170 5 0.0196 10 0.0258 15 0.0321 20 0.0387. 25 0.0465 30 0.0560 35 0.0673 40 0.0816 45 0.0993 50 0.1214 55 0.1477 60 0.1793 65 0.2160 70 0.2564 75 0.2994 80 0.3440 Load Removed 161 SAMPLE 4A T = -4.45°C Yi = 0.905 gm/cc o = 350 psi. Time Deflection (min.) (in.) 0 0.0104 1 0.0166 3 0.0213 5 0.0246 10 0.0317 15 0.0386 20 0.0459 25 0.0542 30 0.0640 35 0.0758 40 0.0898 45 0.1063 50 0.1262 55 0.1494 60 0.1759 65 0.2059 70 0.2390 75 0.2752 80 0.3135 85 0.3526 Load Removed SAMPLE 5A = -4.459C y. = 0.903 gm/cc = 350 psi Time Deflection (min.) (in.) 0 0.0096 1 0.0157 3 0.0202 5 0.0236 10 0.0314 15 0.0390 20 0.0472 25 0.0568 30 0.0681 35 0.0821 40 0.0990 45 0.1196 50 0.1446 55 0.1734 60 0.2060 65 0.2420 70 0.2908 75 0.3210 80 0.3612 Load Removed SAMPLE 6A = -4.45°C y. = 0.900 gm/cc = 249 psi Time Deflection . (min.) (in.) 0 0.0070 1 0.0106 3 0.0135 5 0.0155 10 0.0194 20 0.0259 30 0.0313 40 0.0362 50 0.0410 60 0.0460 70 0.0510 80 0.0566 90 0.0626 Time Deflection (min.) (in.) 100 0.0696 110 0.0775 120 0.0864 130 0.0964 140 ' 0.1077 Load Removed Rest 7.764 x 104 seconds 0 0.1088 1 0.1112 3 0.1140 5 0.1156 10 0.1190 20 0.0236 30 0.1282 40 0.0322 50 0.1364 60 0.1410 70 0.1462 80 0.1520 90 0.1585 100 0.1658 110 0.1745 120 0.1846 130 0.1960 140 0.2008 150 0.2232 153 0.2264 Load Removed t 7.344 x 104 Res seconds 0 0.2228 1 0.2254 3 0.2282 5 0.2299 10 0.2333 20 0.2384 30 0.2424 40 0.2464 50 0.2508 60 0.2554 70 0.2604 80 0.2663 90 0.2734 100 0.2813 110 0.2903 120 0.3003 130 0.3112 Load Removed SAMPLE lB T = -4.45°C Yi = 0.9025 gm/cc o = 100 psi Time Deflection -(min.) (in.) 0 0.0032 1 0.0042 10 0.0052 60 0.0070 180 0.0099 360 0.0111 540 0.0169 720 0.0202 900 0.0234 1080 0.0268 1260 0.0301 1440 0.0333 1620 0.0367 1800 0.0400 1980 0.0434 2160 0.0466 2340 0.0497 2520 0.0530 2700 0.0564 2880 0.0597 3060 0.0631 3240 0.0666 3420 0.0701 3600 0.0738 3780 0.0776 3960 0.0816 4140 0.0856 4320, 0.0898 4500 0.0941 4680 0.0988 4860 0.1035 5040 0.1087 5220 0.1142 5400' 0.1200 5580 0.1258 5760 0.1319 5940 0.1384 6120 0.1450 Time Deflection (min.) (in.) 6300 0.1517 6480 0.1586 6660 0.1650 6840 0.1727 7020 0.1803 7200 0.1888 7380 0.1967 7560 0.2047 7740 0.2128 7920 0.2212 8100 0.2292 8280 0.2328 8460 0.2460 8640 0.2547 8820 0.2630 9000 0.2713 9180 0.2795 9360 0.2876 9540 0.2952 9720 0.3033 9900 0.3110 10080 0.3192 10260 0.3270 10440 0.3348 10620 0.3425 10800 0.3500 10980 0.3570 11160 0.3648 11340 0.3722 11520 0.3797 11700 0.3868 11880 0.3938 12060 0.4003 Load Removed SAMPLE 23 -4.45°C = 0.903 gm/cc = 150 psi Time Deflection (min.) (in.) 0 0.0030 1 0.0043 10 0.0064 20 0.0076 40 0.0099 100 0.0156 200 0.0239 300 0.0315 400 0.0391 500 0.0467 600 0.0549 700 0.0637 800 0.0746 900 0.0877 1000 0.1031 1100 0.1217 1200 0.1415 1300 0.1635 Time Deflection (min.) (in.) 1400 0.1878 1500 0.2128 1600 0.2381 1700 0.2638 1800 0.2903 1900 0.3166 2000 0.3421 2100 0.3661 2200 0.3891 2300 0.4090 2400 0.4320 2500 0.4520 Load Removed 165 SAMPLE 3B T = -4.45°C Yi = 0.906 gm/cc o = 200 psi Time Deflection Time Deflection (min.) (in.) (min.) (in.) 0 0.0050 340 0.1336 0.0076 360 0.1474 5 0.0104 380 0.1624 10 0.0128 400 0.1780 20 0.0166 420 0.1942 40 0.0226 440 0.2116 60 0.0279 ' 460 0.2294 80 0.0328 480 0.2466 100 0.0378 500 0.2642 120 0.0426 520 0.2812 140 0.0479 540 0.2976 160 0.0530 560 0.3141 180 0.0586 580 ' 0.3306 200 0.0651 600 0.3469 220 0.0722 620 0.3629 240 0.0802 640 0.3789 260 0.0886 660 0.3945 280 0.0978 680 0.4098 300 0.1088 688 0.4158 320 0.1206 Load Removed 166 SAMPLE 4B SAMPLE SB T = -4.45°C T = -4.45°C Yi = 0.903 gm/cc Y1 = 0.903 gm/cc g = 250 psi 0 = 250 psi Time Deflection Time Deflection (min.) (in.) (min.) (in.) 0 0.0072 0 0.0060 1 0.0121 1 0.010 5 0.0172 5 0.0145 10 0.0218 10 0.0180 20 0.0290 20 0.0240 30 0.0352 30 0.0290 40 0.0412 40 0.0340 50 0.0453 50 0.0390 60 0.0515 60 0.0440 70 0.0605 70 0.0496 80 0.0684 80 0.0559 90 0.0777 90 0.0627 100 0.0880 100 0.0704 110 0.0990 110 0.0794 120 0.1119 120 0.0892 130 0.1263 130 0.1001 140 0.1418 140 0.1124 150 0.1585 150 0.1265 160 0.1763 160 0.1419 170 0.1948 170 0.1585 180 0.2139 180 0.1763 190 0.2336 190 0.1950 200 0.2532 200 0.2142 210 0.2730 210. 0.2345 220 0.2927 220 0.2550 230 0.3123 230 0.2740 240 0.3319 240 0.2952 250 0.3512 250 0.3162 260 0.3710 260 0.3369 270 0.3892 270 0.3575 280 0.4092 280 0.3780 290 0.4294 290 0.3982 300 0.4497 300 0.4180 Load Removed Load Removed 167 SAMPLE GB SAMPLE 7B T = -4.45°C T = -4.45°C y. = 0.902 gm/cc Yi = 0.902 gm/cc 6 = 300 psi 0 = 300 psi Time Deflection Time Deflection (min.) (in.) (min.) (in.) 0 0.0150-0.022 0 0.0060 1 0.027 1 0.0104 5 0.0322 5 0.0152 10 0.0369 10 0.0192 20 0.0453 20 0.0256 30 0.0540 30 0.0334 40 0.0639 40 0.0423 50 0.0760 50 0.0526 60 0.0905 60 0.0662 70 0.1090 70 0.0829 80 0.1317 80 0.1040 90 0.1590. 90 0.1303 100 0.1912 100 0.1616 110 0.2277 110 0.1976 120 0.2674 120 - 0.2381 130 0.3095 130 0.2813 140 0.3529 140 0.3269 150 0.3992 150 0.3726 160 0.4448 160 0.4183 170 0.4932 170 0.4621 173 ‘0.5085 180 0.5049 Load Removed Load Removed 168 SAMPLE BB SAMPLE 9B T = -4.45°C T = -4.45°C Yi = 0.905 gm/cc Yi = 0.903 gm/cc 0 = 350 psi 0 = 400 psi Time Deflection Time Deflection (min.) (in.) (min.) (in.) 0 0.0078 0 0.0080 1 0.0138 1 0.0150 5 0.0219 3 0.0210 10 0.0293 6 0.0277 15 0.0368 9 0.0345 20 0.0450 12 0.0418 25 0.0549 15 0.0504 30 0.0666 18 0.0608 35 0.0815 21 0.0742 40 0.0994 24 0.0907 45 0.1218 27 0.1112 50 0.1488 30 0.1368 55 0.1801 33 0.1672 60 0.2158 36 0.2021 65 0.2542 39 0.2407 70 0.2950 42 0.2820 75 0.3371 45 0.3265 80 0.3797 48 0.3730 85 0.4210 51 0.4222 90 0.4614 52 0.4396 95 0.5010 Load Removed 96 0.5088 Load Removed 169 SAMPLE 10B T = -4.45°C Y1 = 0.904 gm/cc First Loading 0 = 200 psi Time Deflection Time Deflection (min.) (in.) (min.) (in.) 0 0.0056 Change to Second Loading 0.0086 0 = 300 psi 6 0.0113 696 0.2568 10 0.0126 697 0.2587 20 0.0152 700 0.2671 40 0.0194 703 0.2768 60 ' 0.0232 705 0.2828 100 0.0299 710 0.2992 140 0.0364 715 0.3172 180 0.0428 720 0.3365 220' 0.0495 725 0.3568 260 0.0568 730 0.3772 300 0.0645 735 0.3973 340 0.0741 740 4 0.4166 380 0.0855 745 0.4360 420 0.0988 750 0.4548 460 0.1147 755 0.4736 500 0.1327 760 0.4918 540 0.1525 765 0.5101 580 0.1748 Load Removed 620 0.1994 660 0.2258 696 . ‘ 0.2508 170 SAMPLE 11B T = -4.45°C Y1 = 0.903 gm/cc First Loading 0 = 300 psi Time Deflection Time Deflection (min.) (in.) (min.) (in.) 0 0.0081 860 0.3718 1 0.0128 920 0.3737 5 0.0177 980 0.3757 10 0.0218 1040 0.3775 20 0.0288 1100 0.3793 30 0.0358 1160 0.3812 40 0.0432 1220 0.3831 50 0.0519 1280 0.3850 60 0.0620 1280 0.3850 70 0.0747 1340 0.3871 80 0.0897 1400 0.3892 90 0.1085 1460 0.3910 100 0.1316 1520 0.3932 110 0.1590 1580 0.3954 120 0.1910 1640 0.3977 130 0.2272 1700 0.4000 140 0.2662 1760 0.4020 150 0.3080 1820 0.4046 158 0.3430 1880 0.4070 1940 0.4094 Change to Second Loading 2000 0.4119 118 psi 2060 0.4144 2120 0.4169 0 0.3368 2180 0.4197 5 0.3380 2240 0.4224 10 0.3388 2300 0.4250 20 0.3408 2360 0.4279 40 0.3431 2420 0.4306 60 0.3445 2480 0.4333 80 0.3437 2540 0.4362 100 0.3467 2600 0.4390 120 0.3476 2660 0.4419 200 0.3508 2720 0.4449 260 0.3528 2780 0.4479 320 0.3549 2840 0.4510 380 0.3570 2900 0.4539 440 0.3589 2960 0.4500 500 0.3608 3020 0.4595 560 0.3627 3080 0.4626 620 0.3627 2140 0.4659 680 0.3664 3200 0.4690 740 0.3683 3260 0.4720 800 0.3700 3320 0.4752 171 SAMPLE 11B (continued) SAMPLE 12B T = '4.45°C Time Deflection _ (min.) (in.) Y1 - 0.903 gm/cc c = 250 psi 3223 8:2gig Time Deflection 3500 0.4850 (min-) (1n-> 3560 0.4875 0 0.0080 3680 0.4941 5 0.0158 3740 0.4976 10 0.0194 3800 0.5009 20 0.0250 3860 0.5040 30 0.0300 3920 0.5074 40 0.0347 3980 0.5105 50 0.0395 4040 0.5139 50 0.0446 4100 0.5170 A 70 0.0500 4160 0.5203 Load Removed 80 0.0562 90 0.0628 100 0.0704 110 0.0792 120 0.0892 3130 0.1004 Load Removed Rest 1800 seconds 0 0.0908-0.0986 1 0.1025 5 0.1086 10 0.1147 20 0.1270 30 0.1404 40 0.1556 50 0.1729 50 0.1917 70 0.2118 30 0.2333 90 0.2549 100 0.2779 110 0.3011 120 0.3254 130 0.3485 140 0.3738 150 0.4007 Load Removed SAMPLE 13B T = -4.45°C Y1 = 0.903 gm/cc o = 250 psi. Time Deflection (min.) (in.) 0 0.0086 5 0.0170 10 0.0206 20 0.0263 30. 0.0312 40 0.0364 50 0.0416 60 0.0472 70 0.0536 80 0.0606 90 0.0688 100 0.0783 110 0.0894 119 0.1006 Load Removed Rest 3600 Seconds 0 1 5 10 20 30 40 50 60 70 80 90 100 110 120 130 140 150 160 170 180 190 200 210 220 230 240 0.0910- Load Removed 0.0983 0.1017 0.1088 0.1137 0.1226 0.1319 0.1424 0.1550 0.1699 0.1866 0.2050 0.2247 0.2451 0.2663 0.2880 0.3097 0.3314 0.3531 0.3746 0.3958 0.4167 0.4374 0.4579 0.4785 0.4988 0.5195 0.5400 SAMPLE 143 T = -4.45°C Yi = 0.902 gm/cc o = 250 psi Time Deflection (min.) (in.) 0 0.0068 5 0.0152 10 0.0188 20 0.0246 30 0.0298 40 0.0348 50 0.0401 60 0.0456 70 0.0520 80 0.0591 90 0.0674 100 0.0769 110 0.0878 120 0.1003 Load Removed Rest 5400 Seconds 90 100 110 120 130 140 150 160 170 180 190 200 210 220 230 240 250 260 270 0.0908- Load Removed 0.0975 0.1006 0.1110 0.1166 0.1228 0.1294 0.1372 0.1460 0.1566 0.1689 0.1832 0.1992 0.2167 0.2354 0.2548 0.2747 0.2948 0.3148 0.3348 0.3546 0.3744 0.3934 0.4126 0.4316 0.4500 0.4684 0.4864 0.5044 0.5220 T = —4 45°C Y1 = 0.903 /cc 0 = 250 p51 Time Deflection (min.) (in.) 0 0.0086 5 0.0171 10 0.0205 20 0.0259 30 0.0306 40 0.0356 50 0.0407 60 0.0462 70 0.0522 .80 0.0592 90 0.0671 100 0.0766 110 0.0872 120 0.0997 Load Removed Rest 7200 seconds 0 0.0890-0.0970 1 0.1000 10 0.1088 20 0.1150 30 0.1207 ‘40 0.1265 50 0.1330 60 0.1404 70 0.1494 80 0.1600 90 0.1727 100 0.1869 110 0.2028 120 0.2202 130 0.2388 140 0.2581 150 0.2780 160 0.2979 170 0.3184 180 0.3391 190 0.3598 200 0.3806 210 0.4010 220 0.4215 230 0.4414 240 0.4618 250 0.4822 260 0.5023 Load Removed SAMPLE 16B T = -4.45°C Y1 = 0.904 gm/cc o = 23 psi Time Deflection (hr.) (in.) 0 0.0013 1 0.00234 10 0.00360 20 0.00463 40 0.00620 60 0.00733 80 0.00834 100 0.00918 120 0.00980 140 0.01068 160 0.01147 180 0.01222 200 0.01319 220 0.01408 240 0.01479 260 0.01552 280 0.01612 300 0.01689 320 0.01777 340 0.01858 360 0.01940 380 0.02039 400 0.02121 420 0.02218 440 0.02288 460 0.02358 480 0.02419 500 0.02487 520 0.02509 Load Removed 174 SAMPLE 1H SAMPLE 2H T = -4.45°C (Tensile Creep Test) Y1 = 0.910 gm/cc T = -4.45°C o = 209 psi Y1 = 0.909 gm/cc 02 = 74 psi Time Deflection (min.) (in.) Time Deflection (min.) (in.) 0 0.009 1 0.0117 0 0 10 0.0157 1 0.025 20 0.0188 10 1.39 40 0.0238 20 1.64 60 0.0280 40 1.69 80 0.0321 ,100 1.86 100 0.0362 200 1.97 120 0.0404 340 2.04 140 0.0448 460 2.05 160 0.0496 660 2.27 180 0.0547 720 2.29 200 0.0606 1100 2.64 220 0.0670 1280 2.79 240 0.0742 1460 3.01 260 0.0821 1763 3.28 280 0.0908 2060 3.55 300 0.1070 2630 4.04 320 0.1114 2880 4.22 340 0.1234 3229 4.44 360 0.1355 3522 4.67 380 0.1489 4115 5.16 ,400 0.1629 4370 5.45 420 0.1775 4872 5.87 440 0.1925 5470 6.40 446 0.2084 6273 7.38 480 0.2254 6876 7.95 500 0.2429 7602 8.60 520 0.2614 8350 9.30 540 0.2810 9146 10.30 Load Removed 10578 11.98 11240 12.80 12056 13.82 12982 14.65 14282 16.02 Load Removed 175 SAMPLE 3H Biaxial Creep Test T = -4.45°C ‘ Y1 = 0.911 gm/cc 01 = 192 p31 Time Deflection (min.) (in.) 0 0.0074 10 0.0093 20 0.0106 40 0.0128 60 0.0147 120 0.0194 180 0.0239 240 0.0282 300 0.0330 360 0.0383 420 0.0444 480 0.0518 540 0.0608 600 0.0712 660 0.0841 720 0.0990 780 0.1162 840 0.1358 900 0.1578 960 0.1803 1020 0.2025 1120 0.2380 1180 0.2638 1240 0.2910 1300 0.3192 1360 0.3473 1420 0.3758 03 = 100 psi Time Deflection (min.) (in.) 0 0 5 0.02 10 0.08 15 0.10 20 0.12 30 0.18 40 0.21 50 0.23 60 0.30 80 0.39 100 0.46 140 0.56 200 0.75 402 1.50 465 1.82 580 2.59 700 3.69 730 4.02 760 4.38 830 5.28 1122 9.70 1150 10.20 1214 11.40 1246 12.02 1287 12.78 1321 13.50 1404 15.17 Load Removed 176 SAMPLE 3H: Time, Principal Strains and Octahedral Shear Strains Time 61 e3 82 eoct. (min.) % % 0 0 0 0 0 20 0.0016 0.00072 -0.0023 0.0034 40 0.0027 0.0013 -0.0040 0.0058 100 0.0057 0.0028 -0.0085 0.0121 200 0.0092 0.0045 -0.0137 0.0197 300 0.0129 0.0068 -0.0197 0.0283 400 0.0178 0.0089 -0.0267 0.0385 500 0.0240 0.0102 -0.0342 0.0497 600 0.0327 0.0160 -0.0487 0.0702 700 0.0450 0.0216 -0.0666 0.0962 800 0.0609 0.0280 -0.0889 0.1285 900 0.0788 0.0362 -0.1150 0.1664 1000 0.0993 0.0443 -0.1436 0.2080 1100 0.1200 0.0537 -0.l737 0.2516 1200 0.1400 0.0623 -0.2023 0.2930 1300 0.1597 0.0728 -0.2325 0.3363 1400 0.1800 0.0828 -0.2628 0,3300 “7117117111117:(11111711111114?