Access the full text.
Sign up today, get DeepDyve free for 14 days.
References for this paper are not available at this time. We will be adding them shortly, thank you for your patience.
Rat e an d size effec to s n th e deformatio n an d fractur e o w f ar m an d floatin g columna frreshwater I G En lhar ti A t o Universi Mecha ca E l ng eer g Rate and size effects on the deformation and fracture of warm and floating columnar freshwater ice Iman El Gharamti DOCTORAL DISSER TATIONS ec yt la ma am ni ni in Aalto University publication series DOCTORAL DISSERTATIONS 134/2021 Rate and size effects on the deformation and fracture of warm and floating columnar freshwater ice Iman El Gharamti A doctoral dissertation completed for the degree of Doctor of Science (Technology) to be defended, with the permission of the Aalto University School of Engineering, at a public examination held at the lecture hall 216 of the school on 29 October 2021 at 12:00. Remote connection link: https://aalto.zoom.us/j/66036070494 Aalto University School of Engineering Mechanical Engineering Solid Mechanics L Supervising professor Professor Jukka Tuhkuri, Aalto University, Finland Thesis advisor Professor John P. Dempsey, Clarkson University, USA Preliminary examiners Professor Sveinung Løset, Norwegian University of Science and Technology, Norway Dr. David Cole, Cold Regions Research and Engineering Lab, USA Opponent Professor Sveinung Løset, Norwegian University of Science and Technology, Norway Aalto University publication series DOCTORAL DISSERTATIONS 134/2021 © 2021 Iman El Gharamti ISBN 978-952-64-0533-9 (printed) ISBN 978-952-64-0534-6 (pdf) ISSN 1799-4934 (printed) ISSN 1799-4942 (pdf) http://urn.fi/URN:ISBN:978-952-64-0534-6 Unigrafia Oy Helsinki 2021 Finland Printed matter Printed matter 1234 5678 4041-0619 N Abstract Aalto University, P.O. Box 11000, FI-00076 Aalto www.aalto.fi Author Iman El Gharamti Name of the doctoral dissertation Rate and size effects on the deformation and fracture of warm and floating columnar freshwater ice Publisher School of Engineering Unit Mechanical Engineering Aalto University publication series DOCTORAL DISSERTATIONS 134/2021 Series Field of research Mechanical Engineering Manuscript submitted 25 May 2021 Date of the defence 29 October 2021 Permission for public defence granted (date) 7 September 2021 Language English Monograph Article dissertation Essay dissertation Abstract Climate change has led to warmer and fragmented ice, and thus has increased the interest in understanding and modelling the fracture behavior and creep deformations of warm ice. The thesis explores the fracture and creep response of warm and floating columnar freshwater S2 ice under different loading scenarios, by conducting large scale experiments in the Ice Tank of Aalto University. A program of nineteen mode I fracture tests, using deeply cracked edge-cracked rectangular plates, that varied the test sizes, loading types, and loading rates was conducted. The ice was very warm with a temperature of about -0.3 C at the top surface. The ice was loaded in the direction normal to the columnar grains, and the loading conditions divided the test program into two parts. In the first part, fourteen tests were conducted in displacement control (DC) and loaded with different rates monotonically to fracture. The plates covered a size range of 1:39, the largest for ice tested under laboratory conditions, with three plate sizes: 0.5m x 1m, 3m x 6m and 19.5m x 36m. In the second part, five tests of 3m x 6m plates were loaded in load control (LC) under creep/cyclic-recovery loading and monotonic loading to fracture. For the DC tests, methods for both the linear elastic fracture mechanics (LEFM) and a non-linear viscoelastic fictitious crack model (VFCM) were derived to analyze the data and calculate values for the apparent fracture toughness, crack opening displacement, stress-separation curve, fracture energy, and size of the process zone near a crack tip. Issues of notch sensitivity and minimum size requirements for polycrystalline homogeneity were addressed. Size and rate effects were interrelated as rate dependent size effects and size dependent rate effects. The loading rates applied led to test durations from fewer than 2 seconds to more than 1000 seconds, leading to an elastic response at the highest rates and a viscoelastic response at the lower rates. Under the LC tests' loading conditions, the ice response was overall elastic-viscoplastic; no significant viscoelasticity or major recovery were detected. Moreover, there was no clear effect of the creep loading on the fracture properties at crack growth initiation: the failure load and crack opening displacements. Several factors were discussed as possibly contributing to the observed behavior, and the effect of the very warm ice temperature was highlighted. Schapery's model of nonlinear thermodynamics was tested and validated against the experimental response at the crack mouth. The VFCM and Schapery's model were coupled with the Nelder-Mead's optimization scheme to obtain the constitutive parameters, by matching the displacement records generated by the model and measured by the experiment. Further, different methods for computing the fracture energy were applied, and the values were compared regarding the effects of loading type, rate and scale. Keywords Freshwater ice, fracture, creep, size effect, rate effect, viscoelasticity ISBN (printed) 978-952-64-0533-9 ISBN (pdf) 978-952-64-0534-6 ISSN (printed) 1799-4934 ISSN (pdf) 1799-4942 Location of publisher Helsinki Location of printing Helsinki Year 2021 172 http://urn.fi/URN:ISBN:978-952-64-0534-6 Pages urn … and say: “My Lord! Increase me in knowledge.” … we have no knowledge except what You have taught us. Indeed, it is You who is the Knowing and the Wise. The Holy Quran (20:114 - 2:32) 2 Preface As I write the preface today, I close my eyes and the years begin to move in reverse, slowly ticking backward, rolling back to Jan 4, 2016. Thinking back on that day of my life, I recall everything as clearly as if it were all still unfolding before my eyes. I remember standing at the airport holding two suitcases loaded with food from back home that my Mom packed. My mind vividly replays how I arrived to Finland and what it felt to be alone in cold and dark place. My life during PhD has not been all "rainbows and butterflies", but more or less a roller coast ride. I was riding waves of sadness, happiness, stress, relief, pain, homesickness, success ... all in one ride. There are so many people to thank as I wrap up the writing of my PhD thesis. My first and foremost appreciation goes to my supervisor Prof. Jukka Tuhkuri and my advisor Prof. John Dempsey. Their trust in me since the very beginning granted me the doctoral position and paved my way towards a successful PhD life. Their cooperation and assistance were vital in conducting my research and helping me stay on track. Special thanks go to Jukka for welcoming me in his “Solid Mechanics" group and for granting me the opportunity to travel to Svalbard and work as a fieldwork assistant with other researchers from NTNU (Trondheim, Norway) and UNIS (Svalbard, Norway). Besides the academic guide, Jukka and John have great characters. Jukka is a great listener and supporter; his door was always open for whatever issues I had. John has a great sense of humour; in most of the Skype calls we had, we started with what’s happening around the world and concluded in the last few minutes about research. Working with John, mostly remotely during the whole PhD, proved to me undoubtedly that the success of the work relies mainly on a mutual understanding between the student and the advisor, regardless of whether their offices are in the same building or in different continents. My work during the PhD years was funded though the Finland Distinguished Profes- sor programme "Scaling of Ice Strength: Measurements and Modeling", and through the ARAJÄÄ research project, both funded by Business Finland and the industrial partners Aker Arctic Technology, Arctech Helsinki Shipyard, Arctia Shipping, ABB 3 Preface Marine, Finnish Transport Agency, Suomen Hyötytuuli Oy, and Ponvia Oy. I was also funded by the Doctoral Program of the School of Engineering. The financial support of all my funders is gratefully acknowledged. The great part of my PhD studies was experimental. I would like to sincerely thank people as Prof. Arttu Polojärvi, Kari Kantola, Teemu Päivärinta, Dr. Mikko Suominen, Veijo Laukkanen and every person who provided any kind of help in conducting my experiments at the Aalto Ice Tank. I want to thank Prof. Sven Bossuyt, Dr. Antti Forsström and Kim Widell who took care of the Digital Image Correlation (DIC) work in my experiments. I am thankful to Dr. Otto Puolakka who assisted my work in the laboratory cold room and always had great ideas to share. My gratitude extends to my group members and colleagues (esp. Dr. Hanyang Gong, Anastasia Markou and Dr. Ville-Pekka Lilja) who added a friendly touch to the stressful work atmosphere. I am grateful to Dr. Mingdong Wei who was always ready to help me at subzero temperatures whenever I was stuck with the ice; together we formed the "Panda and Jääkarhu" team. Special thanks go to Dr. David Cole for training me in thin sections techniques and polarized light procedures during his visit to Aalto University. My thankful words reach Profs. Arttu Polojärvi and Luc St-Pierre who assigned me as teaching assistant in their courses. Not to forget to thank Dr. Kari Santaoja for several useful and enlightening discussions. This dissertation could not have been completed without the valuable comments of the pre-examiners Prof. Sveinung Løset and Dr. David Cole. I will be forever indebted to the support of my brother Moha who acted as my numerical advisor; I greatly appreciate his useful advice and help on various numerical issues examined in the thesis. I would like to thank my brother Moustafa for getting me through tough analytical solutions. I am deeply thankful to my aunts and cousins who have tolerated my tense mood during my study years. My love reaches my friends in Finland (Dandoon, Rima, Diza, Nesma and Lubna) who made me feel right at home. Throughout these years, my mother Wafaa has been the buffer between me and the obstacles that were threatening to stop me. It was her support that allowed me to go along with the decision of coming to Finland. Since then, I have been like sailing in a boat in the middle of a sea. Sometimes, the sea is calm. Sometimes, the wind picks up. My mother’s voice, always echoing through my head, was the anchor that prevented my vessel from rocking from side to side and forced me to move forward. Her beautiful face had reassured me, more times than I could ever count, that I was always capable of doing better. You are the greatest gift from God to me and the most beautiful-hearted mother among all the women I met and those that I haven’t. Thank you for providing me with more-than-enough love. 4 Preface My dearest Murtaza, your presence has been my comfort and peace during the PhD days. You are a man who thinks with straight-line logic and weighs a situation without the cloud of worries that affects my own judgement. I have always poured out my concerns to you; and patiently, you were ready to address any surfacing of my negative thoughts and weed out all the what-ifs I bring into the picture. You came into my life, unfroze me from the ice I have been studying, and showed me a new meaning of life that really mattered. A sincere thank you from my heart to a person who is good, all the way to his soul. Finally, I am dedicating this dissertation to the loving memory of my beloved family members (Souad, Jiji and Khalil) who departed this life but continue to mean so much to me. Until we meet again. . . With timeless love. . . Espoo, September 27, 2021, Iman El Gharamti 5 Preface 6 Contents Preface 3 Contents 7 List of Publications 11 Author’s Contribution 13 List of Figures 15 Symbols 25 1. Introduction 29 1.1 Motivation ............................. 29 1.2 Historical Background of Fracture Mechanics .......... 29 1.2.1 Linear Elastic Fracture Mechanics (LEFM) ...... 30 1.2.2 Cohesive Mechanics .................. 32 1.3 Time-Dependent Behavior ..................... 36 1.4 S2 Columnar Freshwater Ice .................... 38 1.5 Fracture Behavior of S2 Columnar Freshwater Ice ........ 39 1.6 Time-dependent Behavior of S2 Columnar Freshwater Ice .... 40 1.7 Objectives, Scope and Structure of the Thesis ............ 41 2. Grown Ice and Experimental Setup [PI and PII] 45 2.1 Experiments ............................ 45 2.1.1 Ice Spraying Trials ................... 46 2.1.2 Ice Growth, Thickness, and Temperature ......... 47 2.1.3 Grain Size and C-axis Orientation ........... 48 2.1.4 Specimen’s Geometry .................. 51 2.1.5 Specimen Preparation and Experimental Procedure . . 52 2.1.6 Fractographic Examinations .............. 55 7 Contents 3. Fracture of Warm S2 Columnar Freshwater ice Under Monotonic Loading: Size and Rate Effects [PI] 59 3.1 Linear Elastic Fracture Mechanics Model ............. 60 3.2 Viscoelastic Fictitious Crack Model ................ 62 3.3 Results ............................... 66 3.3.1 Linear Elastic Fracture Mechanics Analysis ...... 66 3.3.2 Nominal Strength Analysis ............... 69 3.3.3 Viscoelastic Fictitious Crack Model Analysis ...... 71 3.3.4 Notch Sensitivity Analysis ............... 79 3.4 Discussion .............................. 81 4. Fracture of S2 Columnar Freshwater Ice Under Creep/Cyclic-Recovery Loading [PII] 85 4.1 Creep/Cyclic-Recovery Fracture Experiments .......... 86 4.1.1 Specimen ........................ 86 4.1.2 Creep-Recovery and Monotonic Loading Profile . . . 86 4.1.3 Cyclic-Recovery and Monotonic Loading Profile .... 87 Nonlinear Time-Dependent Modeling of S2 Columnar Freshwater Ice 88 4.2 4.3 Results ............................... 92 4.3.1 Experimental and Modelling Results .......... 92 4.3.2 Effect of the Creep and Cyclic Sequences on the Fracture Properties ........................ 92 4.3.3 Ice Response Under the Testing Conditions ...... 94 4.3.4 Nonlinear Modelling Analysis ............. 96 4.4 Discussion ............................. 99 5. Fracture Energy of Columnar Freshwater Ice: Influence of Loading Type, Loading Rate and Size [PIII] 103 5.1 Fracture Energies ......................... 103 Fracture Energy at Crack Growth Initiation via the J- 5.1.1 integral (J ) ....................... 103 5.1.2 Work-of-Fracture (W ) ................. 104 5.1.3 Fracture Energy at Crack Growth Initiation via the VFCM (G ) ......................... 106 VFCM 5.2 Results and Discussion....................... 106 6. Conclusion 111 6.1 Grown Ice and Experimental Setup [PI, PII and PIII] ........ 111 8 Contents 6.2 Fracture of Warm S2 Columnar Freshwater Ice Under Monotonic Loading: Size and Rate Effects [PI] ................ 112 6.2.1 Summary ........................ 112 6.2.2 Research Suggestions .................. 113 6.3 Fracture of S2 Columnar Freshwater Ice Under Creep/Cyclic- Recovery Loading [PII] ...................... 114 6.3.1 Summary ........................ 114 6.3.2 Research Suggestions .................. 115 6.4 Fracture Energy of Columnar Freshwater Ice: Influence of Loading Type, Loading Rate and Size [PIII] ................ 116 6.4.1 Summary ........................ 116 6.4.2 Research Suggestions .................. 116 A. Weight Function For An Edge-Cracked Rectangular Plate 117 Bibliography 119 Publications 127 9 Contents 10 List of Publications This thesis consists of an overview of the following publications which are referred to in the text by their Roman numerals. I Iman E. Gharamti, John P. Dempsey, Arttu Polojärvi and Jukka Tuhkuri. Fracture of warm S2 columnar freshwater ice: size and rate effects. Acta Materialia, 22-34, 202 II Iman E. Gharamti, John P. Dempsey, Arttu Polojärvi and Jukka Tuhkuri. Creep and fracture of warm columnar freshwater ice. The Cryosphere, 2401–2413, 15 2021. III Iman E. Gharamti, John P. Dempsey, Arttu Polojärvi and Jukka Tuhkuri. Fracture energy of columnar freshwater ice: Influence of loading type, loading rate and size. Materialia, 101188, 20 2021. 11 List of Publications 12 Author’s Contribution Publication I: “Fracture of warm S2 columnar freshwater ice: size and rate effects” The author ran the models, analyzed the experimental and model results and wrote the first draft of the paper. The other authors supervised the work and edited the manuscript. All the authors conducted the experiments. Publication II: “Creep and fracture of warm columnar freshwater ice” The author ran the model, analyzed the experimental and model results and wrote the first draft of the paper. The other authors supervised the work and edited the manuscript. All the authors conducted the experiments. Publication III: “Fracture energy of columnar freshwater ice: Influence of loading type, loading rate and size” The author generated and analyzed the results and wrote the first draft of the paper. The other authors supervised the work and edited the manuscript. 13 Author’s Contribution 14 List of Figures 1.1 (a) Kirsch stress distribution around a hole and (b) Inglis stress solution around an ellipse in an infinite plate under uniaxial tension. 30 1.2 Westergaard stress solution around a crack in an infinite plate under equibiaxial tension. The equation mentioned represents the stress along the crack plane. ........................ 31 1.3 linear elastic fracture mechanics (LEFM) stress solution near the tip of a crack under Mode I. K is the stress intensity factor. .... 32 1.4 Dugdale (left) and Barenblatt (right) cohesive models. In Dugdale’s model, the cohesive stress in the process zone (PZ) is constant at the yield strength (σ ) of the material. In Barenblatt model, the cohesive stress (σ(x)) is function of the distance from the traction- free crack tip. ........................... 33 1.5 FCM: (a) idealization of the cohesive crack; (b) stress distribution in the PZ before, and (c) after the development of a FDPZ; (d) stress-separation curve (). δ (2U ) and δ are the crack-opening- v v v displacement and the rate of crack opening, respectively. The subscript v indicates the viscoelastic characterization [PIII]. . . . 34 1.6 Strain responses to a (a) step stress applied between time t and t ; displaying (b) purely elastic, (c) elastic-viscoelastic, (d) elastic- viscoplastic, and (e) elastic-viscoelastic-viscoplastic behavior. The figure was edited as David Cole suggested. Deformation was changed into strain in Figs. 1.6d and 1.6e. The elastic recovery is now added to Fig. 1.6d. ....................... 37 1.7 General strain-time curve showing the three stages (primary, sec- ondary and tertiary) of time-dependent strain. Adapted from San- taoja (1990). ............................. 37 15 List of Figures 1.8 (a) Maxwell, (b) Kelvin-Voigt, and (c) Maxwell-Kelvin mechanical models. E and η are the elasticity and viscosity constants of the spring and dashpot, respectively. ................. 38 1.9 Chemical structure of ice Ih along the basal plane. ........ 38 1.10 Overview of the thesis structure. ................. 43 2.1 Aalto Ice Tank is a unique testing basin in size and shape. .... 45 2.2 Spraying setup consisting of pressure washers and barrels and placed inside the carriage. ..................... 46 2.3 Complete spraying setup attached to the moving carriage. .... 46 2.4 (a) The edge cracked rectangular plate of width H, length L, and crack length A used in the experiments. (b) The crack orientation and direction of crack propagation; also the columnar grains are shown. The dotted rectangular box represents the ice cut for storage and analysis (PI). .......................... 48 2.5 (a) Temperature profile. Each data point represents the average of measurements taken at the same depth of different ice cores throughout the one month duration of the test program. (b) Grain size distribution. Each data point is measured from one thin section (PII). ................................ 49 2.6 (a) Milling machine used to (b) mill the ice and (c) produce thin section. ............................... 50 2.7 Vertical thin sections photographed between crossed polaroids showing columnar freshwater ice (a) at the top of the ice sheet and (b) at the bottom of the ice sheet. The arrows indicate the growth direction. .............................. 50 2.8 Crystalline textures and associated c-axis orientation plots of hori- zontal thin sections photographed between crossed polaroids. The thin sections were taken from the (a) top, (b) middle, and (c) bot- tom of the ice sheet. In the type of axis projection plot used, a horizontal c-axis plots on the circumference and a vertical c-axis would be at the center. Each plot consists of 100 poles of the basal planes measured by a four-axis universal Rigsby stage. ....... 51 2.9 A sketch of the servo-controlled hydraulic loading system. .... 53 2.10 (a) Loading device; inductive displacement transducer (LVDT)s arrangement in a (b) 0.5m x 1m specimen and (c) 3m x 6m speci- men. The digital image correlation (DIC) pattern stamped on the ice surface is shown. ....................... 54 16 List of Figures 2.11 Crack path through the LVDTs in a (a) 0.5mx1m and (b) 3mx6m specimens, as observed directly on the ice surface. ........ 54 2.12 Vertical strain plots computed by DIC technique and showing clearly the crack path for (a) slow 0.5mx1m test (RP6), (b) fast 0.5mx1m test (RP3), (c) slow 3mx6m test (RP13), and (d) fast 3mx6m test (RP11) (see Table 5.1). The arrow indicates the direc- tion of crack propagation. ..................... 55 2.13 Horizontal thin sections showing a crack path (a) from 2 cm behind the initially-sharpened crack tip to 15 cm ahead of it and (b) from 15 cm to 35 cm ahead of the initial crack tip. Both thin sections are from a depth of 24 cm from the top of the ice sheet. The arrows indicate the direction of crack propagation. ............ 56 2.14 (a) Dried replica being peeled off the ice fracture surface and (b) its fractographic examination by scanning electron microscope (SEM). 56 3.1 The edge cracked rectangular plate of width H, length L, and crack length A (PI). ........................... 60 3.2 (a) Illustration of the fictitious crack model. (b) Stress-separation curve describing the behavior in the PZ and showing the control points (see Eq. 3.25) used in the optimization procedure. (c) A flowchart illustrating the forward and inverse problems used for matching the experimental and model results (PI). ........ 63 3.3 (a) Load-CMOD records for the 3m x 6m samples with loading −1 rates as indicated in the figure. The 39.3kPa ms and 1.83 −1 kPa ms data were smoothed by using moving averages (PI). (b) The pre-peak load-CMOD record. ................. 68 3.4 Load-displacement records for specimen RP4 loaded at rate 6.22 √ √ −1 −1 kPa ms (a) and for specimen RP3 loaded at rate 57.1kPa ms (b). Both specimen had the dimensions of 0.5m x 1m. Locations, where crack mouth opening displacement (CMOD), crack opening displacement at about half the length of the crack (COD), and near crack-tip opening displacement (NCOD)1 were measured, are shown in Fig. 3.1. ......................... 68 3.5 Measured apparent fracture toughness as a function of loading rate. First-order power-law fits were applied separately to the data from the larger specimen (3m x 6m and 19.5m x 36m) and from the smaller specimen (0.5m x 1m) (PI). ................ 69 17 List of Figures 3.6 Measured initiation crack opening displacements at the crack mouth (a) and near the crack tip (b) as a function of loading rate. First- order power-law fits were applied to the data. In (b), the power-law fit was calculated separately for the larger specimen (both 3m x 6m and 19.5m x 36m) and from the smaller specimen (0.5m x 1m) (PI). Note that a small error in PI has been fixed here: the power-law exponent in (b) for the small samples should be -0.03 instead of -0.01. ................................ 69 3.7 Nominal tensile strength σ as a function of loading rate (a) and specimen size (b). First-order power-law fits were applied to the data. In (b), the data for the 0.5m x 1m and 3m x 6m specimen is obtained from the power law fits at different loading rates; the −1 loading rate of K = 0.871 kPa ms is indicated with the vertical dashed line in (a). ......................... 70 3.8 Optimization results of RP5. Convergence of (a) the cohesive stresses, (b) the critical crack opening displacement, (c) the creep compliance constant, and (d) the objective function after a certain number of iterations. ........................ 72 3.9 Experimental and model results for RP9 (3m x 6m). (a) Load at the crack mouth, see Fig. 3.1 and Eq. (3.1). (b) Displacement - time records. (c) Load - displacement records. The CMOD data was smoothed with moving average smoothing with a span of 10% successive data points. ....................... 72 3.10 Experimental and model results for RP14 (19.5m x 36m). (a) Load at the crack mouth, see Fig. 3.1 and Eq. (3.1). (b) Displacement - time records. (c) Load - displacement records. (The corresponding figures in PI are edited; (0,0) is added to the model plots.) .... 73 3.11 The crack opening displacement at crack growth initiation at X = A (a) and the actual fracture energy (b) as a function of loading rate. Viscoelastic fictitious crack model was used in the analysis. First-order power-law fits were applied separately to the data for the larger specimen (3m x 6m and 19.5m x 36m) and for the smaller specimen (0.5m x 1m) (PI). .................... 73 3.12 Stress-separation curves under FDPZ conditions for the 0.5m x 1m specimen (a) and for the 3m x 6m and 19.5m x 36m specimen (b) at different loading rates. The number next to each curve reflects the index of the experiment in the Legend (PI). .......... 75 18 List of Figures 3.13 The attained stress-separation curves for the (a) 0.5m x 1m speci- mens and for the (b) 3m x 6m and 19.5m x 36m specimens. The number next to each curve reflects the index of the experiment in the Legend [PIII]. ......................... 75 3.14 (a) Experimental and VFCM results for the crack profiles and (b) PZ profiles by the VFCM for the 3m x 6m and 19.5m x 36m spec- imens. The total separation between the upper and lower crack surfaces is given by 2U ; U is half of δ which is the viscoelastic v v v crack-opening-displacement (see Fig. 1.5). X = A is the coordi- nate of the traction-free tip, and a is the normalized traction-free crack length (a = A /L). The number next to each curve reflects 0 0 the index of the experiment in the Legend [PIII]. .......... 77 3.15 Variation of the process zone size (PZ) with loading rate and speci- men size. First-order power-law fits were applied separately to the data for the larger specimen (3m x 6m and 19.5m x 36m) and for the smaller specimen (0.5m x 1m) (PI). .............. 78 3.16 The growth of the normalized process zone with normalized time for the small, 0.5m x1m, specimens (a) and for the large, 3m x 6m and 19.5m x 36m, specimens (b). The process zone size was normalized by L and the time axis by the time to failure (t )of each experiment (Table 3.1). .................... 78 3.17 Notch sensitivity (σ /σ ) as a function of the normalized crack n t length for the 0.5m x 1m (a), 3m x 6m (b), and 19.5m x 36 m (c) specimens. Lines of constant brittleness numbers are shown. The notch sensitivity of σ /σ = 0.4, 1 are indicated with the horizontal n t dashed lines. The normalized crack length of each specimen size is shown by the vertical dashed line. (d) Notch sensitivity (σ /σ ) n t as a function of the loading rate. First-order power-law fits were applied to the data for the 0.5m x 1m and 3m x 6m specimens (PI). 80 1/2 3.18 Variation of creep compliance (J = 1/E +Ct , given in Eq. 3.17) as a function of loading rate (K) for different specimen sizes (PI). (J in PI was changed to J in the thesis to distinguish between the creep compliance (J ) and the J-integral (J)). ........... 83 4.1 (a) Specimen geometry, edge cracked rectangular plate of length L, width H, and crack length A (PII). ................ 86 4.2 Loading consisting of (a) creep-recovery and (b) cyclic sequences followed by a monotonic fracture ramp. The number above each segment indicates the duration in s (PII). .............. 87 19 List of Figures 4.3 Experimental results for the (a) peak load P , (b) crack mouth max opening displacement CMOD and (c) near crack tip opening dis- placement NCOD1 at crack growth initiation, as a function of time to failure t for the monotonically-loaded DC experiments (Chapter 3) and the creep/cyclic and monotonically-loaded LC experiments (PII). ................................ 93 4.4 Measured load versus CMOD for the (a) displacement control (DC) experiments (Chapter 3), (b) load control (LC) experiments, and (c) LC experiments up to the peak load (PII) ........... 93 4.5 Load versus CMOD over the (a) creep-recovery cycles for RP15 and the (b) cyclic-recovery sequences for RP17. (c) Schematic illustration of the hysteresis load-displacement diagram. The whole of the hysteresis loop area is the energy loss per cycle. The dashed area is the part of that total that is due to the viscoelastic mechanism and the rest is due to viscous processes (PII). ........... 94 4.6 Experimental results for RP16. (a) Load at the crack mouth, see Fig. 4.1. (b) Displacement - time records. (c) Load - displacement record. (d) Typical response of a Maxwell model, consisting of a nonlinear spring and nonlinear dashpot, to a constant load step. (e) CMOD vs time plot showing the first creep sequence. [PII]. . . . 95 4.7 Experimental results for RP17. (a) Load at the crack mouth, see Fig. 4.1. (b) Displacement - time records. (c) Load - displacement record (PII). ............................ 96 4.8 Experimental and model results for RP16. (a) Load at the crack mouth, see Fig. 4.1. (b) CMOD - time records: Schapery’s model with and without viscoelasticity vs the experimental data (PII). . . 97 4.9 Experimental and model results for RP17. (a) Load at the crack mouth, see Fig. 4.1. (b) CMOD - time records (PII). ....... 98 4.10 Contribution of each individual model component to the total CMOD displacement for (a) RP16 and (b) RP17 (PII). ...... 98 5.1 Description of the area used for (a) J (A ) and (b) W (A ) Q J f W Q f calculations. Load versus crack mouth opening displacement for (c) RP4 and (d) RP12, illustrating the A and A areas [PIII]. . 105 J W Q f 5.2 Description of the area used for J (A ) calculations for metals, Q J as defined by the ASTM E1820-20b. Adapted from ASTM E1820- 20b AST (2020). .......................... 105 5.3 Load-CMOD records for the (a) 0.5m x 1m DC (b) 3m x 6m DC, (c) 19.5m x 36m DC and (d) 3m x 6m LC experiments [PIII]. . . . 107 20 List of Figures 5.4 Variation of J and G as a function of the time to failure for Q VFCM the DC experiments [PIII]. ...................... 107 5.5 (a) The apparent fracture energy at crack growth initiation (J ) and (b) the work-of-fracture fracture energy (W ) as a function of the time to failure for the DC experiments. In (a), first-order power-law fits were applied separately to the data for the larger specimen (3m x 6m and 19.5m x 36m) and for the smaller specimen (0.5m x 1m) [PIII]. ................................ 108 5.6 (a) The apparent fracture energy at crack growth initiation (J ) and (b) the work-of-fracture fracture energy (W ) as a function of the time to failure for the 3m x 6m DC and LC experiments. First-order power-law fits were applied to the data for the DC experiments. Note in (a), two LC experiments RP16 and RP17 gave the same J value [PIII]. ............................ 109 5.7 Fracture energy ratio W /J as a function of the time to failure for f Q the DC and LC experiments [PIII]. ................ 110 21 List of Figures 22 List of Acronyms CMOD crack mouth opening displacement COD crack opening displacement at about half the length of the crack CZ cohesive zone CZM cohesize zone model DC displacement control DIC digital image correlation ECRP edge-cracked rectangular plates FCM fictitious crack model LC load control LEFM linear elastic fracture mechanics LVDT inductive displacement transducer N-M Nelder-Mead NCOD near crack-tip opening displacement PZ process zone SEM scanning electron microscope SIF stress intensity factor SSY small scale yielding VFCM viscoelastic fictitious crack model 23 List of Acronyms 24 Symbols A total length of the crack, including the traction-free crack and the process zone A traction-free crack length A area under the load-CMOD curve up to crack growth initiation A area under the entire load-CMOD curve a normalized crack length by the crack-parallel dimension L (A/L) a normalized traction-free crack length by the crack-parallel dimension L (A /L) 0 0 a time scale shift factor in Schapery’s model C creep compliance constant C elastic compliance of Schapery’s model C viscoelastic compliance of Schapery’s model ve C viscoplastic compliance of Schapery’s model vp D length of contact between the ice and the loading device d normalized ice-loading device contact length by the crack-parallel dimension L (D/L) d average grain size av E short-term elastic modulus G fracture energy actually consumed ac G full fracture energy under fully-developed process zone conditions G fracture energy computed via the viscoelastic fictitious crack model, the same VFCM as G ac 25 Symbols g , g , g , g nonlinear functions of the load in Schapery’s model 0 1 2 3 H width of the edge-cracked rectangular plate H weight function h thickness of the ice h normalized weight function J energy release rate given by the J-integral J creep compliance J apparent fracture energy at crack growth initiation via Rice’s J-integral K Mode I stress-intensity factor stress intensity factor (SIF) K Mode I stress-intensity factor due to the cohesive loading coh K critical SIF or fracture toughness IC K apparent fracture toughness K Mode I stress-intensity factor due to the external loading K loading rate L length, crack-parallel dimension, of the edge-cracked rectangular plate P applied force P peak or failure load max t time t time to failure U half of the crack-opening displacement W work-of-fracture X position along the crack x normalized position (X/L) β brittleness number δ crack-opening displacement, equals to 2U δ elastic component of the crack-mouth-opening displacement CMOD 26 Symbols ve δ viscoelastic component of the crack-mouth-opening displacement CMOD vp δ viscoplastic component of the crack-mouth-opening displacement CMOD δ critical crack-opening displacement, equals to 2U ,at X = A given a fully cr cr 0 developed δ viscoelastic crack-opening displacement δ viscoelastic crack-opening displacement due to the external loading vσ δ viscoelastic crack-opening displacement due to the cohesive loading process vcoh zone δ crack-opening displacement at crack growth initiation, at X = A 0 0 δ rate of crack-opening F objective function ν Poisson’s ratio ψ, ψ reduced times of Schapery’s model Σ crack face pressure on 0 ≤ X ≤ D σ normal tensile traction in the cohesive zone coh σ nominal tensile stress at the crack tip σ tensile strength σ /σ notch sensitivity n t σ yield strength 27 Symbols 28 1. Introduction 1.1 Motivation The main goal of this thesis is to gain a better understanding of the fracture behavior of columnar freshwater ice under different loading scenarios. Understanding the deformation and fracture processes of columnar freshwater ice is important in many engineering problems. For example, freshwater ice sheets fracture when in contact with ships, river ice fractures during interaction with bridge piers, and thermal cracks form in lakes and reservoirs. In fact, climate change has led to a warmer, thinner, and broken ice. In addition, the applications like river ice breakup happen in late spring and the ice is very warm. Accordingly, the warming climate increases the importance of studying warm ice, and warming ice increases the importance of creep deformations. But, historically cold ice has been studied typically. It is well-known that the deformation and fracture processes of freshwater ice are highly dependent on temperature, strain rate, sample size, grain type and grain size. Qualitatively, high temperature and low strain rate lead to creep behaviour and ductile fracture; low temperature and high strain rate lead to elastic behaviour and brittle fracture. However, quantitatively these relations are not well known. Generally speaking, unless we restrict our interest on the short time scales where only elastic response is relevant, the creep deformations must be modeled to obtain the true fracture behavior. In materials with time-dependent properties, the fracture and creep deformations are coexistent. 1.2 Historical Background of Fracture Mechanics Fracture mechanics offers a mathematical study of the science of cracks: nucleation, propagation, energy, stress fields, failures, etc. The field started unofficially at the turn of the 20th century with new analytical linear elastic solutions formulated for stresses 29 Introduction at holes in 1898, then at ellipses in 1913. Fracture mechanics research was officially born in 1920 with Griffith’s energy-based analysis of cracks. Later contributions were motivated by several crack-caused failures that happened during World War II and led to an expansion of research in the field. Early research was focused on linear elastic fracture mechanics (LEFM). Nowadays, cohesive mechanics illustrated by the cohesize zone model (CZM) or the fictitious crack model (FCM) is the basic nonlinear approach used to describe the fracture of quasibrittle materials. 1.2.1 Linear Elastic Fracture Mechanics (LEFM) Kirsch (1898) used linear elasticity in his study of stress concentrations and derived a linear elastic solution for stresses around a hole in an infinite plate. His solution contains the well known factor-of-three stress concentration at the hole under uniaxial loading, as shown in Fig. 1.1a. ߪ ߪ ୫ୟ୶ ୶ ͵ߪ ͵ߪ ஶ ஶ ௬௬ ߪ ൌߪ ͳʹ ߪ ஶ (a) (b) Figure 1.1. (a) Kirsch stress distribution around a hole and (b) Inglis stress solution around an ellipse in an infinite plate under uniaxial tension. The second step in the development of LEFM was done by Inglis (1913). Inglis derived a linear elastic solution for stress field surrounding an ellipse in an infinite plate under uniaxial tension, as portrayed in Fig. 1.1b. Inglis result predicts that the maximum stress at the ellipse’s tip goes to infinity as the radius of curvature at the tip goes to zero forming a crack. Griffith (1921) used Inglis linear elastic solution to develop his energy-based failure criterion of cracks. He computed the strain energy release rate associated with crack growth and compared it with the work required to break the atomic bonds ୫ୟ୶ ୫ୟ Introduction and propagate the crack. Griffith’s criterion only applies to brittle materials because it doesn’t account for the energy dissipation associated with plastic deformation near the crack tip. This was later addressed by Irwin (1948) and Orowan (1949), who modified Griffith’s initial development by adding the energy due to plastic deformation at the crack tip. Westergaard (1939) applied an Airy stress function of complex numbers to derive a complete solution for the stress field surrounding a crack in an infinte plate under equibiaxial tension (Fig. 1.2). ߪ ߪ ஶ ஶ ʹܣ ܻൌͲǡܺܣ ߪ ൌߪ ൌ ଶ ͳെሺ ሻ Figure 1.2. Westergaard stress solution around a crack in an infinite plate under equibiaxial tension. The equation mentioned represents the stress along the crack plane. The stress values along the crack plane is given by a simple equation, as shown in Fig. 1.2. However, computing the stress at any other position near the crack requires a heavy work with complex functions. This hinders the general understanding of the stress state surrounding a crack. Irwin (1957) showed that Westergaard’s result could be greatly simplified in the area immediately surrounding the crack tip. Irwin’s approximate solution can replace Westergaard’s exact solution near the crack tip only. Irwin’s expressions lose its accuracy away from the crack tip. The importance of Irwin’s work lies in the fact that his solution led to the definition of the stress intensity factor (SIF), which is the most important parameter of LEFM. LEFM expressions for the crack tip stress field under Mode I are shown compactly in Fig. 1.3, and it is clear that LEFM predicts an unavoidable singularity and infinite stresses at the crack tip for any applied loading. In short, LEFM is a one-parameter fracture model, and it is based on the calculation of a critical stress intensity factor, fracture toughness, which measures a material’s resistance to fracture and is denoted by K . LEFM is limited to situations where IC two main conditions are satisfied: 1) an initial crack exists in the body and 2) the material behavior is linearly elastic and where any nonlinear zone at the crack tip is ୷୷ ୶୶ Introduction ʹܣ ܭ ܭ ߪ ൌ ݂ ߠǢ ߪ ൌ ݃ ߠǢ ʹߨݎ ʹߨݎ ߬ ൌ ݄ሺߠሻ ʹߨݎ Figure 1.3. LEFM stress solution near the tip of a crack under Mode I. K is the stress intensity factor. small compared to other dimensions of the specimen; a condition known as small scale yielding (SSY) (Rice, 1965). 1.2.2 Cohesive Mechanics As previously mentioned, LEFM has proven a useful tool for solving fracture problems provided that two conditions (a pre-existing crack and SSY) are satisfied. However, in a real material, the size of the nonlinear zone, due to creep and plasticity (ductile ma- terials) or microcracking (quasi-brittle materials), doesn’t always satisfy the condition of SSY. Besides, LEFM cannot be used to analyze bodies with no initial crack or with blunt notches [Elices et al., 2002]. Particularly speaking about quasi-brittle materials (such as ice, concrete, rock, and ceramic), the attempts to apply classical LEFM turned out to be misleading [Dempsey et al., 2018], and the fracture toughness results were very scattered and highly dependent on several factors; such as size, loading rate, grain size, temperature, tip acuity, etc. Thus, alternative nonlinear fracture models were needed. The cohesive zone model (CZM) or fictitious crack model (FCM) proved successful to analyze the fracture of concrete, ice and other quasi-brittle materials. A review of the properties of the model was done by many researchers [Bazant and Planas, 1998; Bilby et al., 1963; Smith, 1974; Elices et al., 2002]. A brief presentation of the model development and its main ingredients is presented in the following paragraphs. LEFM claims that all fracture processes take place at the crack tip. On the other ୶୷ ୷୷ ୶୶ Introduction hand, cohesive mechanics eliminates the stress singularity of LEFM by assuming the presence of a cohesive zone (CZ) or process zone (PZ) ahead of the tip where all nonlinearity takes place. The physical crack is a traction-free crack, and the PZ is a fictitious crack whose faces are bridged by unbroken ligaments and are able to transfer cohesive stresses. Dugdale (1960) developed a strip yield model for ductile materials in plane stress. He assumes the existence of a plastic zone at the crack tip. The plastic zone is a PZ whose behavior is represented by a uniform stress distribution equals to the yield strength of the material in a plane stress situation. In Dugdale’s PZ, the cohesive stress is maintained at a constant value irrespective of the crack opening displacement, as shown by the left side of Fig. 1.4. elastic Traction-free ߪሺݔሻ ߪ crack tips ʹܣ PZ PZ Figure 1.4. Dugdale (left) and Barenblatt (right) cohesive models. In Dugdale’s model, the cohesive stress in the PZ is constant at the yield strength (σ ) of the material. In Barenblatt model, the cohesive stress (σ(x)) is function of the distance from the traction-free crack tip. Later, Barenblatt (1962) developed a model for brittle materials analogous to the Dugdale model. However, he assumed that the cohesive stress inside the PZ is not constant, but varying with the distance from the traction-free crack tip, as shown by the right side of Fig. 1.4. Dugdale and Barenblatt models were restricted to the analysis of the fracture processes near the tip of a preexisting crack with the assumption of a small PZ in comparison to other dimensions in the structure. Hillerborg et al. (1976) built on the ideas of the pioneering work of Barenblatt and Dugdale and extended their approaches with two main generalizations. 1) The cohesive crack can develop anywhere in a specimen or a structure, and not only ahead of a preexisting crack tip. In other words, the model applies to all situations of short cracks, long cracks, and no cracks at all. 2) The small size requirement is relaxed, i.e. the process zone size can be large. Hillerborg developed the FCM or CZM. The FCM has proved successful in modeling the fracture behavior of several quasi-brittle materials. The key idea of the FCM is the concentration of all nonlinear fracture mechanisms into a localized process zone, 33 Introduction traction-free fictitious bulk crack crack material A PZ ߜൌʹܷ ߪ ൌ݂ߜሺ ǡ ߜ ሻ ୡ୭୦ ௩ ௩ ߜ ൏ߜ ௩ ߜ ൌߜ ௩ ݂ሺߜ ǡ ߜ ሻ ௩ ௩ ୡ୭୦ ݂ሺߜ ǡ ߜ ሻ ௩ ௩ Figure 1.5. FCM: (a) idealization of the cohesive crack; (b) stress distribution in the PZ before, and (c) after the development of a FDPZ; (d) stress-separation curve (). δ (2U ) and δ are the v v v crack-opening-displacement and the rate of crack opening, respectively. The subscript v indicates the viscoelastic characterization [PIII]. 34 Introduction located on the crack line, as shown in Fig. 1.5a. The length of a traction free crack is denoted as A and the total length of a crack including the process zone as A, giving a length of the process zone as PZ = A − A . The process zone initiates at a point where the maximum principal stress reaches the tensile strength (σ ) (Fig. 1.5a). The process zone develops with its surfaces being attracted to each other by cohesive stresses. The cohesive tensile stress (σ ) carried at a particular point within the PZ coh over the distance (A − A ) is a function of the separation of the PZ at that point. The points lying within the PZ first transmit the full tensile strength, but then the stress transmitted softens with crack opening (Fig. 1.5b). The total separation between the upper and lower crack surfaces is denoted by δ = 2U; U is half of the crack opening. At some critical separation (δ , Fig. 1.5c), no tensile stress can be transmitted and the PZ is termed a fully developed process zone (FDPZ). The behavior inside the process zone is described by a constitutive relation, known as the stress-separation law (σ − δ law), which relates the softening stresses to the increasing crack opening displacements (Fig. 1.5d). The FCM assumes that there is a smooth closure of the surfaces within the cohesive zone, i.e. dδ/dX = 0at X = A (known as the closure condition). The main elements that characterize the cohesive model are shown in Fig. 1.5: the length of the process zone (PZ), the maximum cohesive stress (σ ) that the material coh in the process zone can transfer and which initiates the growth of the process zone (tensile strength, σ ), the critical displacement beyond which the material is fully separated and can transfer no more stresses (δ ), the shape of the σ − δ curve, and the fracture energy represented by the area under the σ − δ curve. Precisely speaking, Fig. 1.5d shows a σ−δ curve obtained under conditions in which a fully-developed-process-zone (FDPZ) has formed. The full fracture energy (G )is consumed. A FDPZ is achieved when two conditions are satisfied at X = A : σ = σ 0 coh t and δ = δ . Based on the testing conditions (load control versus displacement control, 0 c ice temperature, operating deformation mechanisms, etc.), either a FDPZ or a non- FDPZ is obtained before fracture occurs. In the case of a non-FDPZ, the cohesive stress at X = A does not drop to zero and the fracture occurs at δ <δ ; thus, only a 0 0 c portion of the σ − δ curve / G (Fig. 1.5d) is attained. Several approaches have been proposed to quantify the softening law and can be categorized into: direct (experimental) and indirect (experimental-numerical). A pure experimental method would involve stable tensile tests to determine the parameters of the σ − δ curve. However, this approach has proven difficult and few reliable results have been obtained. As summarized by Elices et al. (2002) and Van Mier and Van Vliet (2002), several technical, material, and environmental problems arise and prevent the achievement of a stable tensile test. Some technical difficulties include: 1) 35 Introduction achieving the perfect alignment of the specimen in a system with very low eccentricity, 2) solving the gripping problem such that failure occurs far away from the platens, and 3) selecting a stiff and stable test-control system. Other material and specimen related problems arise from 1) the heterogeneity of the material disrupting the uniform stress distribution across the specimen’s cross section and causing multiple cracking and 2) internal rotations of the specimen itself leading to asymmetric modes of fracture and nonuniform crack opening across the specimen. Environmental conditions of the specimen such as the temperature gradients can also cause deviation from uniform stress distribution. For these reasons, direct methods are replaced with indirect methods, typically through a combined experimental-numerical procedure. Indirect methods have been employed both in concrete and ice research and proved promising (Elices et al., 2002; Mulmule and Dempsey, 1997). These techniques apply inverse analysis in an iterative procedure to obtain the softening curve of the FCM. 1.3 Time-Dependent Behavior The material’s behavior is highly dependent on the loading scenario: loading rate, loading type (monotonic, creep, cyclic), and testing conditions (displacement vs load control). In fact, every scenario produces a new response of the same material. Fig. 1.6 displays different strain responses to a step stress applied between time t and t 0 1 (Fig. 1.6a). Fig. 1.6b shows a purely elastic response, where the material behavior is independent of time. The material reacts instantaneously to the load application and removal. However, Figs. 1.6c, 1.6d, and 1.6e show a time-dependent behavior where the material response is not instantaneous but delayed with time. In these cases, the constitutive relation is given by the stress and the strain rate (instead of strain). General time-dependent deformation (Fig. 1.6e) consists of three components: elastic, viscoelastic (delayed elastic, anelastic), and viscoplastic (viscous). Elastic deformation is the instantaneous, time-independent and recoverable deformation. Vis- coelastic deformation is the time-dependent and recoverable deformation. Viscoplastic deformation is the time-dependent and permanent (unrecoverable) deformation. Costa Andrade (1910) developed the idea of dividing the time-dependent deforma- tion curve into three main stages. Each stage is characterized by different deformation mechanisms. Division into stages allows to describe each stage separately and for- mulate an expression for each dominant deformation. Fig. 1.7 shows a general creep strain-time curve made up of four main stages: 1) Instantaneous stage is characterized by instantaneous purely elastic and recovered deformation (ε ), 2) primary (transient) stage is characterized by a decreasing strain rate and a time-dependent viscoelastic and ve recoverable deformation (ε ), 3) secondary (steady state) stage is characterized by a Introduction Figure 1.6. Strain responses to a (a) step stress applied between time t and t ; displaying (b) purely elas- 0 1 tic, (c) elastic-viscoelastic, (d) elastic-viscoplastic, and (e) elastic-viscoelastic-viscoplastic behavior. The figure was edited as David Cole suggested. Deformation was changed into strain in Figs. 1.6d and 1.6e. The elastic recovery is now added to Fig. 1.6d. constant strain rate and a time-dependent viscoplastic and unrecoverable deformation vp (ε ), and 4) tertiary stage is characterized by an increasing strain rate and accelerating te deformation (ε ) leading to complete failure of the material. Failure ௩ ௩ 2 3 4 Primary Secondary Tertiary Figure 1.7. General strain-time curve showing the three stages (primary, secondary and tertiary) of time-dependent strain. Adapted from Santaoja (1990). Creep behavior is represented by mechanical models of springs and dashpots, mod- elling the elastic and time-dependent behavior, respectively. The two simplest models are the Maxwell and Kelvin-Voigt models which consist of a spring and dashpot in series and parallel, respectively, as shown in Figs. 1.8a and 1.8b. The other general- ized models (for example, Fig. 1.8c) are constructed by considering more and more elements. The more elements the model has, the more accurate it will be in describing ௧ Introduction ܧ ߟ (a) (b) (c) Figure 1.8. (a) Maxwell, (b) Kelvin-Voigt, and (c) Maxwell-Kelvin mechanical models. E andη are the elasticity and viscosity constants of the spring and dashpot, respectively. the real response of real materials. 1.4 S2 Columnar Freshwater Ice Ice exists in a number of forms, but only the ordinary ice Ih is present in a significant amount in nature. In ice Ih, the oxygen atoms occupy the points of a hexagonal lattice in which each oxygen atom is tetrahedrally coordinated with four other oxygen atoms. Each oxygen atom is associated with two hydrogen atoms, as shown in Fig. 1.9. The oxygen molecules are most closely packed on basal planes, situated perpendicular to the principal hexagonal axis, called the c-axis. The c-axis also coincides with the optical axis of each ice crystal. Figure 1.9. Chemical structure of ice Ih along the basal plane. Freshwater ice possesses important optical properties: birefringence, uniaxial (one optical axis), and transparent (especially when bubble-free). Birefringence means that the ice exhibits the property of double refraction when it is examined under polarized light. The interference colors under cross-polaroids depend on the orientation of the crystals and allows individual grains in polycrystalline ice to be identified. This property is crucial in the analysis of thin sections to investigate the crystalline texture, c-axis orientation, and the ice type. The ice type examined in this thesis is columnar freshwater S2 ice. As defined by Michel and Ramseier (1971), S2 ice is columnar polycrystalline ice with a random and horizontal c-axis orientation in most of its grains. S2 ice individual crystals exhibit the 38 Introduction same color but with different intensities under cross-polarized light [Shestov, 2018]. 1.5 Fracture Behavior of S2 Columnar Freshwater Ice Generally speaking, freshwater ice is a quasi-brittle material that tends to be fairly brittle and fails by microcracking instead of yielding. However, the exact failure mode depends critically on the temperature and loading conditions, including grain size, thickness, loading rate, crack orientation, loading direction, notch acuity, loading configuration, and specimen size. Investigations of their effects have been studied by Liu and Loop (1972), Goodman and Tabor (1978), Liu and Miller (1979), Hamza and Muggeridge (1979), Urabe and Yoshitake (1981), Timco and Frederking (1982), Danilenko (1985), Parsons and Snellen (1985), Timco and Frederking (1986), Nixon and Schulson (1986), Tuhkuri (1987), Bentley et al. (1988), Dempsey et al. (1989), Parsons et al. (1989), Wei et al. (1991), Weber and Nixon (1996a,b), and Xu et al. (2004). More than five decades ago, Weeks and Assur (1969) stated that "We feel that an understanding of the scale effect in ice testing is essential before a thorough scientific basis can be developed for the utilization of small-scale testing in engineering design problems". This requires a large range in specimen size to be tested in order to reveal the scale effect on the tensile fracture characteristics. However, most of the earlier studies with freshwater ice have used specimens with dimensions of tens of centimeters, and large scale in-situ fracture experiments examining the scaling problem with freshwater S2 ice have not been conducted before. Dempsey (1991) provided a summary of the experimental work on fracture of freshwater ice and suggested that many laboratory specimens may have been too small in terms of the requirement of LEFM. It is important to note that a scale effect may not occur unless the specimen sizes are large enough (Dempsey et al., 1999a). In addition, the general focus of the previous work with freshwater ice has been to characterize the mode I fracture behavior in terms of fracture toughness using LEFM. Most of these toughness studies relied heavily on the principles of LEFM, without ensuring that the necessary requirements have been satisfied (Dempsey et al., 1992). By conducting large-scale in-situ fracture experiments with sea ice, Dempsey et al. (Mulmule and Dempsey, 1999; Dempsey et al., 2018) eventually questioned the applicability of a one parameter fracture mechanics for sea ice, ultimately favoring a viscoelastic fictitious crack model. These, and other, observations led to the suggestion that the fracture of small and large ice specimen may not be similar, and that the applicability of LEFM to even freshwater ice may be limited. At the core of this discussion is the requirement of LEFM that the material behaviour is linearly elastic 39 Introduction except in a small area near a crack tip (Dempsey et al., 2018; Dempsey, 1991). Qualitatively, it can be argued that these conditions prevail when the loading rate is high and the specimen is large. Quantitatively, we know little as to what is a high rate, how large is a large specimen, and how these two questions are affected by temperature, size and type of the ice (granular or columnar), and other parameters. While the viscoelastic fictitious crack model had proved useful for first-year sea ice (Dempsey et al., 2018), it is unknown if it can model the fracture of columnar freshwater ice. In addition, the stress-separation curve for this type of ice has not yet been determined. Another important concern relates to the requirements (number of grains and spec- imen size) needed to ensure notch sensitivity and homogeneous response of the polycrystal. For a specimen to be notch sensitive, the crack and uncracked ligament must both be long enough for the specimens to fail by fracture (Dempsey et al., 1992). The dependence of notch sensitivity on specimen size was discussed by (Carpinteri, 1982) for concrete fracture. However, the combined effects of size and rate on notch sensitivity have not been discussed before. Polycrystalline homogeneity requires that the crack length and the uncracked ligament are significantly larger than the grain size, and the specimen contains sufficient number of grains to be regarded as homo- geneous (Dempsey, 1991; Abdel-Tawab and Rodin, 1993, 1998). As discussed by Dempsey et al. (1999b), polycrystallinity is an issue with large grained ice. Mulmule and Dempsey (2000) speculated that sample homogeneity for first-year sea ice would be obtained if the crack-parallel specimen size harboured at least 200 d , where d av av is the average grain size. For the tensile and compressive testing of un-notched ice cylinders, it is recommended that the cylinder diameter be 15 to 20 times d (Schwarz av et al., 1981). This highlights the huge difference posed by the testing of cracked and uncracked test samples. As soon as one is testing a precracked test sample, the requisite specimen size is measured not in the 10’s of grain size, but the 100’s of the grain size. The requisite test size for polycrystalline homogeneity and notch sensitivity for large grained and warm columnar freshwater ice, considering size and rate effects, is unknown. 1.6 Time-dependent Behavior of S2 Columnar Freshwater Ice As discussed earlier, freshwater ice can exhibit a strong time-dependent behavior; and accordingly, a general constitutive model should incorporate elastic, viscoelastic, and viscoplastic components (Jellinek and Brill, 1956; Sinha, 1978). The importance of each component depends on the problem studied. For example, thermal deformations of ice in dams can have a time scale of a few days and viscous behaviour dominates. 40 Introduction In ice-structure interaction problems, the time scale of interest is often seconds and hours, so all three components of deformation need to be modeled. The time-dependent behavior of freshwater ice has been addressed with great atten- tion, and several constitutive models were developed (Michel, 1978a; Sinha, 1978; Le Gac and Duval, 1980; Ashby and Duval, 1985; Sunder and Wu, 1989; Mellor and Cole, 1983; Cole, 1990; Duval et al., 1991; Sunder and Wu, 1990; Abdel-Tawab and Rodin, 1997; Santaoja, 1990). Constitutive laws can be phenomenological or micromechanical. Micromechanical modeling in ice faces challenges because the characterization of the microscopic mechanisms of ice deformation is still inadequate (Abdel-Tawab and Rodin, 1997). Phenomenological laws are classified into two groups. The first group are empirical- based relations (Sinha, 1978; Schapery, 1969). Their equations relate macroscopic variables: stress/load, strain/displacement, and time. They do not contain state vari- ables that describe the internal state of the material. The functions in these models can be easily calibrated to simulate the experiments. The second group of phenomeno- logical models starts from physically-based models involving internal state variables (dislocation density, internal stresses reflecting hardening, etc ...); they develop dif- ferential equations for the evolution of these variables with time and quantify the dependence of these variables on stress, temperature and strain (Le Gac and Duval, 1980; Sunder and Wu, 1989, 1990; Abdel-Tawab and Rodin, 1997). These models provide insights into the microscopic mechanisms taking place, and the state vari- ables describe the deformation resistance offered by changes in the microstructure of the material. However, they require a proper identification of the deformation mechanisms. The effect of time-dependent loading on the strength of freshwater ice has been examined in the literature. Subjecting freshwater ice to cyclic loading apparently leads to a significant increase in the tensile, compressive, and flexural strength of that ice (Murdza et al., 2020; Iliescu et al., 2017; Iliescu and Schulson, 2002; Cole, 1990; Jorgen and Picu, 1998). On the other hand, no detailed investigation of the effect of creep and cyclic loading on the fracture properties of freshwater ice has been conducted in the past. 1.7 Objectives, Scope and Structure of the Thesis The thesis explores the fracture and creep response of warm columnar freshwater S2 ice under different loading scenarios under laboratory conditions. Large scale experiments were conducted in the Ice Tank of Aalto University. The work is directly relevant to a number of practical problems of freshwater ice (Ashton, 1986), but has 41 Introduction also general relevance in ice research by studying the coupled creep and fracture in a quasi-brittle material. Two main factors make the conducted study unique. The studied size range (1:39, largest size 19.5m x 36m) is the largest to any material under laboratory conditions. Another factor is the very warm temperature of the ice (-0.3 C at the top surface). As discussed, the behavior of warm ice is inevitably complicated by the presence of creep, especially for slower loading rates. The tensile fracture behavior of warm columnar freshwater ice is still a relatively unexplored topic, especially regarding the influence of loading type, loading rate, and specimen size. Sections 1.5 and 1.6 presented several limitations and unanswered questions in earlier research of columnar freshwater ice. To fill the afore-mentioned gaps and to obtain a better understanding of the mechan- ical behavior of columnar freshwater ice, the thesis addresses the following research questions: • RQ1. What are the effects of plate size and loading rate on the fracture behavior of columnar freshwater ice? Are these effects independent or interrelated? No previous study of this ice examined both effects simultaneously or their interrelation. (Chapter 3) • RQ2. Is LEFM a valid model to analyze the fracture data of columnar freshwater ice? (Chapter 3) • RQ3. Is viscoelastic fictitious crack model (VFCM) applicable to study the fracture of columnar freshwater ice? (Chapter 3) • RQ4. What are the requirements and the requisite test size for notch sensitivity and polycrystalline homogeneity for large grained columnar freshwater ice? (Chapter 3) • RQ5. What is the influence of time-dependent loading (creep/cyclic-recovery) on the fracture properties i.e. the failure (peak) load and the crack opening displacements at crack growth initiation? (Chapter 4) • RQ6. How do the testing conditions affect the creep stages (primary/transient and steady-state/secondary) present in the ice? What is the extent to which the elastic, viscoelastic and viscoplastic components contribute to the ice deformation? (Chapter 4) 42 Introduction • RQ7. What is the influence of loading type (monotonic, creep and cyclic), loading rate, and specimen size on the fracture energy of columnar freshwater S2 ice? (Chapter 5) In order to answer these questions, a program of nineteen mode I fracture exper- iments, using deeply cracked edge-cracked rectangular plates (ECRP), that varied the test sizes, loading types, and loading rates was conducted. The ice thickness was homogeneous throughout the parent ice sheet, but it grew from 340 mm to 410 mm during the test program. The ice was very warm; the ice surface was slightly cooler than -0.3 C while the bottom surface, the ice-water interface, manifested an ice temperature of 0 C . Each initial notch was fabricated and tip-sharpened to be 70− 75% of the crack-parallel dimension. The ice was loaded in the direction normal to the columnar grains. The test program was divided into two parts. In the first part, fourteen experiments were conducted in displacement control (DC) and loaded with different rates monotonically to fracture. The loading rates applied led to loading durations from fewer than 2 seconds to more than 1000 seconds. The plates covered a size range of 1:39 with three plate sizes: 0.5m x 1m, 3m x 6m and 19.5m x 36m. In the second part, five experiments of 3m x 6m ECRP were loaded in load control (LC) under creep/cyclic-recovery loading and monotonic loading to fracture. Creep and cyclic sequences were applied maintaining peak loads well below the failure loads, followed by monotonic ramps leading to complete fracture of each specimen. Chapter 3 Chapter 2 {RQ1, RQ2, RQ3, RQ4} Monotonic loading Ice growth/properties Loading rate effect Size effect Microstructural/ LEFM vs VFCM Fractographic Notch sensitivity characterization Lar Larg ge-scale in-situ e-scale in-situ Polycrystalline mode I fracture mode I fracture Experimental setup homogeneity experiments of experiments of warm warm columnar columnar Chapter 4 {RQ5, RQ6} freshwater ice freshwater ice Chapter 5 Creep/cyclic- {RQ7} Effect on fracture recovery loading properties Influence of loading type, loading rate, and test size Deformation Creep stages on the fracture energy components Figure 1.10. Overview of the thesis structure. The rest of the thesis is structured as shown in Fig. 1.10. Chapter 2 presents descriptions of the ice growth, ice properties, microstructural and fractographic char- acterization, and the experimental setup. Chapter 3 discusses the first part of the test 43 Introduction program: the DC monotonic experiments. Linear elastic fracture mechanics is derived and applied as a one-parameter fracture mechanics approach to derive the expressions of the apparent fracture toughness and the crack opening displacement. Then, the chapter introduces a non-linear viscoelastic fictitious crack model and explains how it is used to analyze the experiments. The experimental and model results with the notch sensitivity and homogeneity analysis are reported and analyzed. Chapter 4 discusses the second part of the test program: the LC creep/cyclic-recovery experiments. The ice response is examined under the testing conditions, and Schapery’s model is used to analyze the experimental data. Chapter 5 uses the results of Chapters 3 and 4 to quantify the fracture energy, of all the experiments using different methods, and discuss its sensitivity to loading rate, loading type and plate size. The last chapter summarizes the main findings of the thesis and speculates on some possible future research. 44 2. Grown Ice and Experimental Setup [PI and PII] This chapter presents detailed descriptions of the ice growth, the ice properties, the microstructural and fractographic analysis, and the experimental setup of the test program. The chapter has been adapted from the author’s publications: PI and PII. 2.1 Experiments Large-scale in-situ experiments were conducted in the Aalto Ice Tank, a 40 m x 40 m, 2.8 m deep water basin equipped with a cooling system and a moving measuring carriage (Fig. 2.1). While field work with ice provides results from a natural material in the real environment, laboratory studies allow a control of test parameters and ice texture not possible in the field. Cooling fans Carriage Grown ice Figure 2.1. Aalto Ice Tank is a unique testing basin in size and shape. 45 Grown Ice and Experimental Setup [PI and PII] 2.1.1 Ice Spraying Trials Before the actual experiments, several ice spraying trials were carried out during two months. The spraying was conducted from the measurement carriage. The spraying system consisted of two pressure washers with a maximum pressure of 145 bar and a water flow of 500 liters per hour. Both pressure washers were connected to four water barrels, 200 liters each, with a linking hub, as shown in Fig. 2.2. Carriage Pressure washer Linking hub Barrel Figure 2.2. Spraying setup consisting of pressure washers and barrels and placed inside the carriage. The two pressure washers were connected with rods to spraying nozzles and attached to the carriage, which runs along the length of the basin, as shown in Fig. 2.3 . The rods and the spraying heads were wrapped with foam-covered heating cables to prevent freezing. Spraying rods and head wrapped into Nozzle foam Pushing Scrapers plates Cutting Fishing rods line of the connecting the boom scrapers to the pushing plates Booms Figure 2.3. Complete spraying setup attached to the moving carriage. Basic guidance for growing S2 ice was provided by Gow et al. (1988). The spraying 46 Grown Ice and Experimental Setup [PI and PII] trials helped to understand the conditions needed to generate columnar S2 ice. Some of these factors are: the 1) temperature of sprayed water, 2) volume of sprayed water, and 3) spraying speed. The monitoring of these three parameters helps to control the resultant ice type and grain size. The large-grained ice that forms without spraying should be cleared away before the real spraying. This ice was pushed away using some pushing plates and installed booms (Fig. 2.3), and scrapers were used to cut the ice loose from the boom. Besides, enough water should be sprayed into the air to prevent the formation of open water spots which resulted in large grains’ areas. To do so, the spraying speed was reduced to increase the number of areas where the spraying units overlap; this generated areas of small grains. Moreover, the water droplets sprayed into the air should be fine and cold enough to freeze in the air before touching the water surface. This requires the basin’s water and air temperature to be low enough. The sprayed water droplets were sprayed high enough to stay longer in the air. The spraying nozzles were approximately one meter above the water surface, and the pressure was high enough to push the sprayed water close to the roof. This prevented the remaining of open water areas and allowed the ice to nucleate once the frozen droplet hits the surface. Furthermore, the water in the basin should be as calm as possible without any current or circulation. This was achieved by stopping the air bubbling systems of the basin some time before the spraying started. 2.1.2 Ice Growth, Thickness, and Temperature A sheet of freshwater S2 ice was grown. Protective piping with warming cables was installed along the walls of the tank to protect the ice and the tank itself from any damage during the ice growth. The growth process was initiated by lowering the air temperature in the ice tank to −14 C and bubbling air into the water to obtain a uniform water temperature. When a water temperature of 0.2 C was reached, the air bubbling system was shut off and after approximately 15 minutes a fine mist of water droplets was sprayed into the air above the basin. Once the droplets reached the water surface, they acted as nuclei for ice crystals to form. The temperature of the sprayed water was around 2 C and around 1200 liters of water was sprayed into the air. After four weeks at −14 C, the initial seeded ice layer had developed into a homogeneous sheet of bubble-free columnar S2 ice with a thickness of approximately 34 cm. For the experiments, the ambient temperature was raised to −2 C and maintained at that temperature. The ice thickness was homogeneous throughout the sheet. But, as the test program lasted about a month, the ice sheet grew from the 34 cm to 41 cm during the testing duration. The ice thickness at the crack tip was measured before each test and is reported in Table 3.1. Cutting all the specimens from a single 47 Grown Ice and Experimental Setup [PI and PII] H = 2L Dimensions in mm NCOD3 60-100 NCOD2 NCOD1 COD A = 0.7-0.75 L 0.5A D = 150 CMOD 50-80 (a) (b) Figure 2.4. (a) The edge cracked rectangular plate of width H, length L, and crack length A used in the experiments. (b) The crack orientation and direction of crack propagation; also the columnar grains are shown. The dotted rectangular box represents the ice cut for storage and analysis (PI). homogeneous ice sheet ensured the same grain size and structure for all the specimens. Except specimen size, loading rate, and the gradually increasing thickness, all ice and test parameters were constant. After completing each experiment, a through-the-thickness block of ice was cut from both sides of the crack path as illustrated by the dotted rectangular box in Fig. 2.4b. These blocks extended from 5 cm behind up to 15 − 35 cm ahead of an initial crack tip and allowed the measurement of thickness profiles along the crack paths: the ice thickness varied only a few millimeters along each crack path. The ice blocks studied were clear and transparent; there were practically no air bubbles. The ice blocks were stored in triple plastic bags to prevent sublimation and placed in freezers operating at -20 C Care was taken to mark the correct orientation of the blocks. The vertical temperature profile of the ice sheet was measured daily, at different locations in the ice sheet. Fig. 2.5a shows the temperature profile throughout the whole test program. Each data point represents the average of several measurements taken at the same depth of different ice cores. The ice temperature varied more or less linearly with depth but did not vary much during the test program. It is worth emphasizing that this ice temperature, −0.3 C at the top surface, is very warm in comparison to the majority of laboratory tests performed to date. 2.1.3 Grain Size and C-axis Orientation The microstructure of the ice was analysed by making and studying thin sections in a cold room. The conventional way of thinning ice with a microtome was replaced by a milling machine (Figs. 2.6a and 2.6b). The milling machine proved to be successful in producing a good quality surface finish (2.6c), saving the long working hours with a microtome, and generating larger thin sections than with a microtome. This helped to visualize the complete crack path. In this method, ice sections were frozen on a 48 Grown Ice and Experimental Setup [PI and PII] 0 0 -5 -10 -10 -15 -20 -20 -25 -30 -30 -35 -40 -40 02468 10 -0.4 -0.3 -0.2 -0.1 0 (a) (b) Figure 2.5. (a) Temperature profile. Each data point represents the average of measurements taken at the same depth of different ice cores throughout the one month duration of the test program. (b) Grain size distribution. Each data point is measured from one thin section (PII). plexiglass plate (2.6c), instead of a glass plate. The plexiglass used has as good optical properties as glass, if not better, and provide safer operations. The ice froze fast to the plexiglass and formed a strong bond. With the milling machine, the thickness of the ice sections was reduced to about 0.8 mm, as recommended by Shestov (2018). The thin sections were then examined and photographed under cross-polarized light. Figs. 2.7 and 2.8 show vertical and horizontal sections, respectively. The columnar structure of the ice, with the diameter of the grains increasing with depth, is evident. Several methods can be used to estimate the grain size of ice (Cole, 1986). In the present work, grain size was estimated by using the uniform-sphere-assumption method which relies on estimating the number of grains within a unit area and as- suming that the grains are of uniform size and spherical shape. Fig. 2.5b shows the diameter of the columnar grains in horizontal planes at different depths. The grain size varied between 3 mm at the top and 10 mm at the bottom portion of the ice sheet. The mean grain diameter (d ) was 6.5 mm. av An examination of the c-axis orientation was carried out by analyzing hundreds of grains at different locations and depths along the ice sheet. The crystalline texture and c-axis orientation plots from horizontal thin sections are shown in Fig. 2.8. Each Shmidt net in Fig. 2.8 consists of 100 poles of the basal planes measured with a four-axis universal Rigsby stage from the corresponding horizontal sections - top, middle, and bottom - randomly taken from the ice sheet. Refraction corrections, following Kamb (1962), were applied for the universal stage measurements. In the type of axis projection plot introduced by Langway (1958), a horizontal c-axis would be on the circumference and a vertical c-axis would be at the center. It can be seen that the c-axes of the columnar grains were randomly horizontal. The ice sheet had the same type of textural features throughout the depth and for the whole ice sheet: 49 Grown Ice and Experimental Setup [PI and PII] (a) (b) (c) Figure 2.6. (a) Milling machine used to (b) mill the ice and (c) produce thin section. 10 mm 10 mm (a) (b) Figure 2.7. Vertical thin sections photographed between crossed polaroids showing columnar freshwater ice (a) at the top of the ice sheet and (b) at the bottom of the ice sheet. The arrows indicate the growth direction. 50 Grown Ice and Experimental Setup [PI and PII] the ice was columnar S2 ice. 10 mm 10 mm 10 mm Figure 2.8. Crystalline textures and associated c-axis orientation plots of horizontal thin sections pho- tographed between crossed polaroids. The thin sections were taken from the (a) top, (b) middle, and (c) bottom of the ice sheet. In the type of axis projection plot used, a horizontal c-axis plots on the circumference and a vertical c-axis would be at the center. Each plot consists of 100 poles of the basal planes measured by a four-axis universal Rigsby stage. 2.1.4 Specimen’s Geometry The edge-cracked rectangular plate (ECRP) specimen was used to study the fracture of freshwater ice (Fig. 2.4a). Dempsey and Mu (2014) outlined the theoretical 51 Grown Ice and Experimental Setup [PI and PII] development of the weight function for this geometry with width to length (H/L) ratios of (Fig. 2.4a): 0.25, 0.5,1,1.5, 2, and 4. The weight function is a method that uses the knowledge of a two-dimensional elastic crack solution (the reference solution for any reference crack face pressure) to determine the expressions of the stress intensity factor and crack opening displacement for the same body under arbitrary loading. A wider plate with H/L = 2 was selected for the current study because it offers a higher potential to obtain straight crack path stability and make the crack propagate in a self-similar manner (Lu et al., 2015). The geometry was self-equilibriating; the specimen was completely free-floating and no supports at the edges were required. 2.1.5 Specimen Preparation and Experimental Procedure The mode I fracture tests with loading at the crack mouth were conducted with different loading rates and three specimen sizes: 36m x 19.5m, 6m x 3m, and 1m x 0.5m. The samples covered a size range of 1 : 39 which is the largest employed in laboratory ice experiments. The loading rates applied led to loading durations from fewer than 2 seconds to more than 1000 seconds. The ECRP specimens were cut from the parent ice sheet with an electric chain saw. The cuts were made sufficiently large to prevent fast refreezing between the specimen and the parent ice sheet, and the kerfs around each specimen were kept clear of ice by using shovels. Cracks were cut with the electric chain saw giving a crack of width 8 mm. Directly before each test a sharp crack tip was produced using either a carpet knife with a blade width of 0.71 mm, or a hand saw with a blade width of 1.21 mm intended to be used for lightweight concrete. The length of the sharpened crack was A ≈ 0.7 − 0.75 L, where L is the crack-parallel side length. The final sharpening can help to nucleate microcracks along the crack front, thus effectively creating a very sharp and more realistic crack tip. No significant blunting of the tip was visually observed during the test. The plane of the crack was perpendicular to the plane of the ice sheet; the crack front was vertical as shown in Fig. 2.4b. The ice was stored as a floating parent ice sheet until cut into a specimen, which was subsequently instrumented and tested, before the next specimen was cut out of the parent ice sheet. A closed-loop servo-controlled hydraulic loading device (Fig. 2.9) was installed in a rectangular slot cut at the crack mouth to apply pressure on the crack faces (Figs. 2.4a and 2.10a). The length of the contact between ice and the loading device (D = 150 mm, Figs. 2.4a) was the same for each specimen size and was chosen to keep the contact pressure below 0.5 MPa to avoid crushing of the crack faces. The contact pressure was estimated before the experiments by Equation (3.8) introduced in Chapter 3. Two hydraulic cylinders with maximum forces of 25 kN and 39 kN in compression and two load cells with maximum capacities of 25 kN and 50 kN were used. The 52 Grown Ice and Experimental Setup [PI and PII] feedback displacement-control measurement system control signal feedback coming from the measurement load-control system servo control connected to data-measurement servo system pressure under plate for servo connected to the valve (manifold) cable pressure unit outlet Force Transducer displacement transducer Figure 2.9. A sketch of the servo-controlled hydraulic loading system. smaller cylinder and load cell were used for the smallest sized specimens. Direct measurement of the crack opening displacements was carried out using LVDTs positioned at six locations (Figs. 2.4a, 2.10b and 2.10c): at the centerline of the loading device (the control displacement), at the crack mouth (CMOD), at about half the length of the crack (COD), 10 cm behind the initial sharpened crack tip (NCOD1), 6 − 10 cm ahead of the tip (NCOD2), and 20 cm ahead of it (NCOD3). NCOD denotes near-crack-tip opening displacement. LVDTs with a measuring range of 0 − 2 mm and a resolution of ±3 μm were used. In addition, parallel LVDTs with a larger range were used as secondary sources of data and as backup. No major discrepancies between the primary, high resolution transducers and the secondary transducers were observed. All the LVDTs were mounted on 14.5 mm diameter wooden sticks that were frozen into holes in the ice, on both sides of the crack. The positions of the LVDTs, including the elevation above the ice surface, were kept constant for all the tests. The data were sampled at 1000 Hz. The crack propagation was monitored using the Digital Image Correlation (DIC) method, which is a whole-field measurement technique. The region in the vicinity of the initially sharpened tip was stamped with optimized pattern with a feature size of 3-5 pixels. Generation of the pattern was based on the methodology discussed by Bossuyt (2013). The objective was to generate a pattern with sharp features and high contrast. Taking advantage of the fact that freshwater ice is transparent, a thin hinge plate piston rod amplifier hoses pipes hinge amplifier cylinder plate Grown Ice and Experimental Setup [PI and PII] (a) (b) (c) Figure 2.10. (a) Loading device; LVDTs arrangement in a (b) 0.5m x 1m specimen and (c) 3m x 6m specimen. The DIC pattern stamped on the ice surface is shown. layer of white titanium oxide mixed into water was applied to the ice surface as the base layer, and black carbon powder mixed into water was used to stamp the black features of the pattern, as shown in Figs. 2.10c and 2.11. High speed cameras, placed normal to the deformation surface, have been used to record the images of the deformation surface. Frame rate was selected at 300 frames per second. (a) (b) Figure 2.11. Crack path through the LVDTs in a (a) 0.5mx1m and (b) 3mx6m specimens, as observed directly on the ice surface. In all the tests, the cracks propagated more or less straight, approximately along the x-axis and through the gauges NCOD2 and NCOD3, as shown in Figs. 2.4 and 2.11. Regardless of the specimen size and loading rate, straight cracking path characterized the fracture, as clearly illustrated by the vertical strain (perpendicular to the X-axis, Fig.2.4a) plots in Fig. 2.12, as computed by DIC on top of the ice sheet for different tests’ conditions (Ahmad, 2019). This confirms that the aspect ratio of H/L = 2 used in the tests supports a straight crack propagation. 54 Grown Ice and Experimental Setup [PI and PII] (a) (b) (c) (d) Figure 2.12. Vertical strain plots computed by DIC technique and showing clearly the crack path for (a) slow 0.5mx1m test (RP6), (b) fast 0.5mx1m test (RP3), (c) slow 3mx6m test (RP13), and (d) fast 3mx6m test (RP11) (see Table 5.1). The arrow indicates the direction of crack propagation. It is worth noting that the choice of the gauge setup (Fig. 2.4a) is linked to the modelling procedures and analysis in Chapters 3 and 4. The CMOD, COD and NCOD1 records are important for the implemented VFCM (Chapter 3) and Schapery’s model (Chapter 4). The matching of the measured records with the models’ results leads to the back-calculation of the models’ parameters. The NCOD1, NCOD2 and NCOD3 records are important for the validation of the DIC measurements in the vicinity of the tip. These experiments were the first attempt to apply DIC to study the fracture of ice. Unfortunately, the trial was not very successful. The DIC nicely showed the growth of the crack; however, the resolution of the DIC system was not high enough to measure reliably the deformation. 2.1.6 Fractographic Examinations The ice samples cut from both sides of the crack path were used to study the crack propagation by making horizontal thin-sections as sketched in Fig. 2.4b. Some of the ice samples were cut immediately after an experiment, others were cut one or two days after an experiment. As a result, the crack healed in some samples, and in other samples, the right and left sides of the crack were unattached. It was easy to make thin sections with a healed crack. For unhealed cracks, the two halves of the specimen were matched, and the matched sections were analyzed. The thin sections covered almost the whole crack path and provided numerous grains to study if a crack propagated through a grain or along a grain boundary. It is obvious from Fig. 2.13 55 Grown Ice and Experimental Setup [PI and PII] 10 mm 10 mm (a) (b) Figure 2.13. Horizontal thin sections showing a crack path (a) from 2 cm behind the initially-sharpened crack tip to 15 cm ahead of it and (b) from 15 cm to 35 cm ahead of the initial crack tip. Both thin sections are from a depth of 24 cm from the top of the ice sheet. The arrows indicate the direction of crack propagation. (a) (b) Figure 2.14. (a) Dried replica being peeled off the ice fracture surface and (b) its fractographic examina- tion by SEM. 56 Grown Ice and Experimental Setup [PI and PII] that a transgranular fracture was dominant: the cracks propagated through the grain for almost all the grains. It is worth mentioning that the author did several trials to replicate the fracture surface. Struers silicon solution, made of silicon and hardener, was dropped on the selected areas of the fracture surface, as shown in Fig. 2.14a. After storing the ice sample in the freezer, the solution dried, and the resulting plastic film replica was peeled off ready for examination. Unfortunately, when the dry replica was observed with scanning electron microscope (SEM), no visible fracture features were detected. However, signs of frozen water droplets were observed. Accordingly, this replica procedure of the fracture surface does not work with wet fracture experiments. Although some of the ice was saved immediately after each experiment, the water from the tank melted the fracture features. Water droplets were frozen to the fracture surface, preventing any further analysis of the fracture surface. 57 Grown Ice and Experimental Setup [PI and PII] 58 3. Fracture of Warm S2 Columnar Freshwater ice Under Monotonic Loading: Size and Rate Effects [PI] The following chapter covers the first part of the experimental program. Fourteen displacement control (DC) experiments were conducted to study the fracture behavior of columnar freshwater S2 ice under monotonic loading scenarios. The experiments used three ECRP sizes: 19.5m x 36m, 3m x 6m, and 0.5m x 1m (Fig. 2.4). The crack-parallel dimension covered a size range of 1:39, the largest for ice tested under laboratory conditions. The loading was applied in the direction normal to the columnar grains. The loading rates applied led to test durations from fewer than 2 seconds to more than 1000 seconds (Table 3.1). Two analytical approaches were implemented: a one-parameter linear elastic fracture mechanics approach and a multi-parameter nonlinear viscoelastic approach. These methods were derived to analyze the data and calculate values for the apparent fracture toughness, crack opening displacement, notch sensitivity, stress-separation curve, fracture energy, and size of the process zone near a crack tip. The chapter addresses the research questions RQ1, RQ2, RQ3, and RQ4 (Section 1.7). Specifically, it investigates: 1) the effects of size (scale) and loading rate on the fracture behavior of columnar freshwater ice, 2) the validity of LEFM as a one-parameter fracture tool, 3) the applicability of VFCM to analyze the fracture data, and 4) the requirements for polycrystalline homogeneity and notch sensitivity. The chapter is structured as follows. First, linear elastic fracture mechanics is introduced to derive the expressions of the apparent fracture toughness and the crack opening displacement. Second, a non-linear viscoelastic fictitious crack model is explained and coupled with an optimization model. Finally, the experimental and modelling results are presented and discussed. The chapter has been adapted from the author’s publication (PI). 59 Fracture of Warm S2 Columnar Freshwater ice Under Monotonic Loading: Size and Rate Effects [PI] H = 2L Dimensions in mm NCOD3 60-100 NCOD2 NCOD1 COD A = 0.7-0.75 L 0.5A D = 150 CMOD 50-80 P P Figure 3.1. The edge cracked rectangular plate of width H, length L, and crack length A (PI). 3.1 Linear Elastic Fracture Mechanics Model Expressions for stress intensity factor and crack opening displacement for the edge loaded, edge cracked rectangular plate used in the present experiments (Fig. 3.1) are derived below by following the weight function approach outined by Dempsey and Mu (2014) and reviewed in Appendix A. The crack face pressure Σ(X) by the loading device is ⎪ 0 < X ≤ D Dh Σ(X) = (3.1) 0 D ≤ X < A where P is the applied force, D the length of contact between ice and the loading device, h the ice thickness, X the position along the crack, and A the crack length. The stress intensity factor K(A )is K(A ) = Σ(X)H (A , X)dX (3.2) 0 r 0 where H (A , X) is the weight function (valid for X ≤ A ) given as r 0 0 i− 1 A X H (A , X) = √ G 1 − (3.3) r 0 i L A 2πA i=1 G (i = 1,2,...,5) are functions given by Eq. (A3) in Appendix A and L is the length of the rectangular plate (Fig. 3.1). Inserting Eqs. (3.1) and (3.3) into Eq. (3.2) leads to K(A ) = H (A , X)dX (3.4) 0 r 0 Dh where 2A H (A , X)dX = Z (d, a ), r 0 1 0 G (a ) d i 0 i− with Z (d, a ) = 1 − 1 − , 1 0 2i − 1 a i=1 D A d = , a = (3.5) L L 60 Fracture of Warm S2 Columnar Freshwater ice Under Monotonic Loading: Size and Rate Effects [PI] Thus, P 2A K(A ) = Z (d, a ) (3.6) 0 1 0 Dh π If the critical stress intensity factor (the fracture toughness) is specimen size inde- pendent and specimen geometry independent, it can be denoted by K . Because IC of the lack of standards for determining the proper specimen size, loading rate, etc. for the fracture toughness testing of ice, Dempsey (1991) suggested the notation of apparent fracture toughness K instead of K . Then K(A ) = K when P = P Q IC 0 Q max and it follows from Eq. (3.6) that P 2a max 0 K = Z (d, a ) (3.7) Q 1 0 dh L The maximum contact pressure p = P /Dh follows as max max 2a K 0 p = √ (3.8) max Z (d, a ) 1 0 For the crack opening displacement at a position X along a crack with length A , Dempsey and Mu (2014) gave an expression δ(A , X) = K(T)H (T, X)dT (3.9) 0 r where E is Young’s modulus. For X < D < A , this equation can be put into form D A E δ(A , X) = 2 K(T)H (T, X)dT + K(T)H (T, X)dT (3.10) 0 r r X D An expression for K(T) when 0 ≤ T ≤ D, obtained by using Eq. (8) of Dempsey and Mu (2014), is P T K(T) = Tπ f (η) with η = (3.11) Dh L where f (η) is the dimensionless factor given by Eq. (5) in Dempsey and Mu (2014). The expression for K(T) when D ≤ T ≤ A can be obtained from Eq. (3.6) by replacing A with T and a by η. The expression for COD then reads 0 0 2P 2T π E δ = √ f (η)H (T, X)dT r r Dh π X 2 2T + Z (d,η)H (T, X)dT (3.12) 1 r When 0 < X < D: d a 2P π E δ = √ f (η)Z (x,η)dη + Z (d,η)Z (x,η)dη (3.13) r 2 1 2 πdh 2 x d 61 Fracture of Warm S2 Columnar Freshwater ice Under Monotonic Loading: Size and Rate Effects [PI] in which 5 3 i− X x 2 x = , while Z (x,η) = G (η) 1 − (3.14) 2 i L η i=1 Similarly, when D < X < A : ⎪ A ⎪ ⎪ ⎪ 2P ⎨ 2T ⎬ E δ = Z (d,η)H (T, X)dT (3.15) ⎪ 1 r ⎪ ⎪ ⎪ ⎩ ⎭ Dh π and finally 2P E δ = Z (d,η)Z (x,η)dη (3.16) 1 2 πdh 3.2 Viscoelastic Fictitious Crack Model The fictitious crack model (FCM) introduced in Section 1.2.2 is another modelling approach to analyze the experiments. According to this model, the length of a traction free crack is denoted as A and the total length of a crack including the process zone as A, giving a length of the process zone as PZ = A − A (Fig. 3.2a). The crack-opening displacement at X = A is denoted as δ . The behavior inside the process zone is 0 0 described by a constitutive relation, known as the stress-separation law (σ − δ law), as shown in Fig. 3.2b. The response outside the process zone is governed by an applicable constitutive relation of the bulk material. In the present work, the viscoelas- tic fictitious crack model (VFCM) formulated by Mulmule and Dempsey (1997) is adopted to model the response of the freshwater fracture experiments conducted. The VFCM couples the FCM with the assumption that the behavior of the bulk material is viscoelastic. However, for the freshwater ice fracture experiments conducted, neither the stress-separation law nor the viscoelastic material parameters are known before the experiments, but need to be back-calculated by requiring that the results from the experiments and from the modeling match. An inverse analysis is applied in an iterative procedure to obtain the softening curve of the FCM. The method is outlined below. In this study, the creep compliance J is used to characterize the viscoelastic behavior of the bulk material: 1/2 J (t) = 1/E + Ct (3.17) where E is the short-term elastic modulus measured at the crack mouth, C is the creep compliance constant, and t is the time. Eq. (3.17) is based on laboratory-scale experiments conducted by (Cole, 1993; Schapery, 1993) on saline ice at -10 C under low compressive stress (0.5 MPa). Here, E is obtained from the initial linear portion 62 Fracture of Warm S2 Columnar Freshwater ice Under Monotonic Loading: Size and Rate Effects [PI] (a) (c) traction-free fictitious bulk Choose C, ߪെߜ curve, and PZ size crack crack material Forward Solve the A PZ 0 problem VFCM (Eqs. 3.17-3.24) Obtain CMOD, COD, NCOD1, and PZ records Iterations continue until the solution converges (b) Solve an optimization Inverse scheme by minimizing problem between the model and data (Eqs. 3.25-3.26) Obtain C and ߪെߜ curve Figure 3.2. (a) Illustration of the fictitious crack model. (b) Stress-separation curve describing the behavior in the PZ and showing the control points (see Eq. 3.25) used in the optimization procedure. (c) A flowchart illustrating the forward and inverse problems used for matching the experimental and model results (PI). of the load - crack mouth opening displacement plot using Eq. 3.16 and C through the iterative procedure described below. According to the spirit of the FCM, the growth of the process zone is governed by the interaction of an externally applied load and the cohesive stress such that the resultant stress intensity factor vanishes at the fictitious crack tip: K = K (σ, a, t) + K (σ , a, t) = 0 (3.18) σ coh coh where K, K , and K represent the resultant stress intensity factor (SIF) at the σ coh fictitious crack tip, the SIF due to external loading and the SIF due to the cohesive loading, respectively. The normalized length of the resultant crack is denoted by a = A/L, Figs. 3.2a and 3.1), and t is the time. Using the definition in Eq. (3.2), K and K become coh K (σ, a, t) = L σ(x, t) h (a, x) dx σ r K (σ , a, t) = L σ (x, a, t) h (a, x) dx, a ≥ a (3.19) coh coh coh r 0 where h (a, x) is a normalized weight function, h (a, x) = LH (A, X). The viscoelas- r r r tic crack opening displacement δ is obtained similarly by superposition: δ (x, a, t) = δ (σ, x, a, t) + δ (σ (δ , δ , x), x, a, t) (3.20) v vσ v coh coh v v where δ and δ represent the viscoelastic crack opening displacement due to the vσ v coh external loading, and the viscoelastic crack opening displacement due to the cohesive loading, respectively. The normalized position along the crack is denoted by x = X/L. 63 Fracture of Warm S2 Columnar Freshwater ice Under Monotonic Loading: Size and Rate Effects [PI] The functional dependencies in Eqs. 3.18 and 3.20 are shown to clarify the nonlinearity of the problem. Using Eq. (3.9), the weight function method gives t e dδ δ (x, a, t) = J (t − τ) (x, a(τ),τ) dτ (3.21) v c dτ where J is the creep compliance given by Eq. (3.17), and e e e δ (x, a,τ) = δ (σ, x, a,τ) + δ (δ , x, a,τ) (3.22) v vσ v coh with a(τ) δ (σ, x, a,τ) = 2 L K (s,τ)h (s, x)ds σ r vσ a(τ) δ (δ , x, a,τ) = 2 L K (s,τ)h (s, x)ds (3.23) v coh r v coh max(a ,x) Details of the time integrations are explained in Mulmule and Dempsey (1997). The relation between the cohesive stress, the viscoelastic crack opening displace- ment, and the rate of the crack opening displacement is given by the stress-separation curve, shown in Fig. 3.2b, as follows, f (δ , δ )0 ≤ δ <δ ⎨ v v v c σ (δ , δ ) = (3.24) coh v v ⎪ 0 δ ≥ δ v c where δ and δ represent the critical crack opening displacement and the rate of crack c v opening, respectively. To complete the viscoelastic fictitious crack model for freshwater ice, an iterative procedure was implemented. The availability of the load history and the displacement histories at different positions along the crack provides a way to conduct an inverse analysis and back-calculate the stress-separation (σ − δ) law, the value of the creep compliance constant (C), and the length of the process zone (PZ) (Figs. 3.2). At each iteration, two problems are solved, forward and inverse. The forward problem refers to the VFCM outlined above (Eqs. 3.17-3.24), and the inverse problem corresponds to the optimization procedure described hereinafter. In the forward problem, values for σ−δ curve, C, and PZ are initially assumed – and later obtained though optimization – and the viscoelastic fictitious crack model (Eqs. 3.17-3.24) is solved to obtain the crack opening profile and the length of the process zone. The measured load - time record is applied in discretized steps into the model. The traction-free crack length is known and is discretized into fixed number of stations (5, 22 and 138 for the 0.5m x 1m, 3m x 6m and 19.5m x 36m specimens, respectively). However, the process zone mesh is updated with each load step because the process zone size is increasing. In each load step, new stations are added in the process 64 Fracture of Warm S2 Columnar Freshwater ice Under Monotonic Loading: Size and Rate Effects [PI] zone to keep the mesh fine while keeping the previous stations at constant locations. The governing equations are satisfied at these stations. In each iteration, the crack opening displacements are computed at all the stations. This provides displacement results at different positions along the crack, including the process zone, through an experiment. The length of the process zone is obtained from the requirement that the stress intensity factor is zero at the fictitious crack tip (Eq. 3.18). The solution of the VFCM is considered converged when the crack opening displacements don’t change by more than 0.2% at any of the stations. In the inverse problem, the crack opening profile is assumed to be known, and an optimization scheme using the Nelder-Mead (N-M) algorithm (Nelder and Mead, 1965) is used to obtain the constitutive parameters (C and the σ−δ law) by minimizing the difference between the numerical data and the experimental data. Once the σ − δ law and C are solved, they are used as input in the forward problem and the iteration continues until the best attainable agreement is obtained. A flowchart illustrating the implemented procedure is shown in Fig. 3.2c. In the optimization scheme, five control pairs, shown in Fig. 3.2b, are used to define the σ − δ curve: ⎧ ⎫ ⎪ ⎪ ⎪ ⎪ ⎪ ⎪ δ = 0,δ ,δ ,δ ,δ ⎨ 1 2 3 c ⎬ (3.25) ⎪ ⎪ ⎪ ⎪ ⎪ ⎪ ⎩ ⎭ σ = σ ,σ ,σ ,σ , 0 t 1 2 3 in which δ , δ , and δ are assumed such that δ <δ <δ <δ , and the N-M 1 2 3 1 2 3 c scheme optimizes the control points, δ , σ , σ , σ , and σ . The σ − δ law was c t 1 2 3 back-calculated without any assumption about its shape. The optimization scheme minimizes an objective function F given by the norm of a residual N=45 F = arg min M (δ ,σ ,σ ,σ ,σ , C) − D (3.26) i c t 1 2 3 i δ ,σ ,... c t i=1 where M and D refer to the displacement values given by the model and obtained i i by using the experimental data, respectively. . is the Euclidean norm of a vector. The displacement values that are matched total to 45: CMOD, COD, and NCOD1 (Fig. 3.1) at 15 time instances during the loading. As no LVDT was placed at the initially sharpened crack tip, it was decided to match the displacement values given by NCOD1. 15 instances were chosen to capture the overall shape of the measured load- time record, so that the load given to the model is very similar to the experimental one. Similarly, 5 pairs were selected to capture the overall shape of the stress-separation curve. Choosing more or less points is possible; it is a trade-off between accuracy vs computational complexity. In the current problem, the objective function F is numerical, and accordingly, no closed form solution for the gradient of the objective function is available. Derivative-based optimization methods such as Newton solvers 65 Fracture of Warm S2 Columnar Freshwater ice Under Monotonic Loading: Size and Rate Effects [PI] cannot be used. This is the main motive behind the selection of the Nelder-Mead (N-M) method, which is a derivative-free optimization method (Nelder and Mead, 1965). It starts with an initial guess of C and σ−δ curve. It constructs a simplex of N+1 vertices; N is the number of parameters to be optimized (σ ,σ ,σ ,σ ,δ , and C). In each t 1 2 3 c iteration, N-M computes the value of the objective function at each vertex, removes the vertex with the largest function value, and replaces it with a new vertex with smaller residual by following one of four schemes. The four schemes are reflection, expansion, outside contraction, and inside contraction. This algorithm applies an iterative update of the simplex in shape and size until convergence is reached. The N-M algorithm uses a termination criterion based both on the absolute size of the simplex with respect to the initial simplex and the difference of the function values in the simplex. The optimization problem has constraints: the cohesive stresses are tensile and have a softening behavior (Fig. 3.2b). Besides, upper and lower bounds were provided for the arguments. The bounds for the cohesive stresses were based on the tensile strength of columnar freshwater S2 ice (Carter, 1971; Gold, 1977; Michel, 1978b), for the critical crack opening displacement on the measured displacements at NCOD1 and NCOD2 in each experiment, and for the creep compliance constant on sea ice fracture studies performed by Dempsey et al. (2018). This optimization problem is typically called a least-squares problem when using the Euclidean norm (Eq. 3.26). It is a convex problem because F is a convex function and the feasible set is convex. Thus, the optimization algorithm will converge to the global optimal solution. 3.3 Results 3.3.1 Linear Elastic Fracture Mechanics Analysis Table 3.1 gives the dimensions of the ice samples together with the measured and computed results. The apparent fracture toughness (K ) was computed from the failure load and dimensions by using Eq. (3.7). The loading rate (K) was computed by dividing K by the time to failure (t ). The time to failure varied from about 2 Q f seconds to about 1000 seconds (17 minutes) giving a loading rate range of 0.18 to 57 −1 kPa ms . The elastic moduli (E ) and (E ) were computed by Eq. (3.16) CMOD COD from the initial linear portion of the load - CMOD displacement records and load - COD displacement records, respectively. Some of the E values are missing, caused COD by the fact that the initial portion of the associated load-COD curve was very noisy. No significant difference in the moduli was found. Displacements at the crack mouth 66 Fracture of Warm S2 Columnar Freshwater ice Under Monotonic Loading: Size and Rate Effects [PI] Table 3.1. Specimen dimensions, measured data, and results computed using linear elastic fracture mechanics. ˙ ˙ ˙ Experiment LH A hE E P t σ K K CMOD CMOD NCOD1 NCOD1 0 CMOD COD max f n Q √ √ −1 −1 −1 m m m cm GPa GPa kN s MPa kPa mkPa ms μm μms μm μms RP1 0.5 1 0.35 35 6.9 - 1.7 54.9 0.486 126.2 2.298 33.1 0.602 5.2 0.094 RP2 0.5 1 0.35 35 6.9 - 2.3 51.9 0.657 168.6 3.247 43.3 0.834 7.5 0.145 RP3 0.5 1 0.35 37.1 6.4 - 1.6 1.9 0.431 108.4 57.058 29.1 15.321 7.8 4.128 a a RP4 0.5 1 0.35 37.6 7.0 - 1.9 20.8 0.505 129.2 6.223 37.6 1.811 5.2 0.25 a a RP5 0.5 1 0.35 40.2 7.3 - 2.1 571.7 0.522 132.2 0.231 36.1 0.063 5.4 0.009 RP6 0.5 1 0.35 41.1 5.6 - 2.4 811.4 0.584 145 0.179 59.5 0.073 10.4 0.013 RP7 3 6 2.1 34.5 6.0 6.2 4.6 86.9 0.259 159.4 1.834 236 2.716 47.2 0.543 RP8 3 6 2.1 34.5 6.6 - 3.0 2.7 0.169 106 39.259 157 58.148 22.9 8.482 RP9 3 6 2.1 34.5 6.1 6.2 5.2 148.0 0.293 180.6 1.221 266 1.798 42.8 0.289 RP10 3 6 2.1 34.5 6.7 6.7 4.8 222.2 0.271 166.8 0.751 215.6 0.970 34.6 0.156 RP11 3 6 2.1 36 7.4 7.7 4.2 15.3 0.227 139.4 9.091 174.4 11.374 25.2 1.644 RP12 3 6 2.1 37.2 5.7 5.9 6.8 701.8 0.355 221.3 0.315 370.5 0.528 68.5 0.098 RP13 3 6 2.1 37.6 4.7 4.8 5.7 1027.1 0.295 184.3 0.179 420.3 0.409 63.4 0.062 RP14 19.5 36 14.6 35 5.7 5.7 10 213.5 0.127 186.0 0.871 961.7 4.505 37.2 0.174 Errors in Table 1 of PI have been fixed. and near the crack tip at the crack growth initiation, CMOD and NCOD1 respectively, ˙ ˙ were measured at locations shown in Figure 3.1. CMOD and NCOD1 indicate the displacement rates and were obtained by dividing CMOD and NCOD1 by the time to failure. Attempts were made to measure displacements ahead of the crack also (Fig. 3.1), but those displacements were very small and the measurements failed especially for small specimens and fast experiments and are not analysed here. Fig. 3.3 shows the load-CMOD records for the 3m x 6m specimens. A decrease of the peak load with increase in the loading rate is observed. The pre-peak records (Fig. 3.3b) show an approximately linear load-displacement relation up to the peak load at high loading rates, but at low loading rates, a non-linear relation is observed. Fig. 3.4 shows the load-CMOD data for two 0.5m x 1m specimen: for a low rate in Fig. 3.4a (Experiment RP4) and for a high rate in Fig. 3.4b (Experiment RP3). For RP3, a period of crack growth took place after the peak load before complete fracture happened: the load-displacement plots show portions of approximately constant load and increasing displacement. The post-peak crack growth was observed for experiment RP3 only, all the other experiments fractured completely at the peak load. Fig. 3.5 shows the apparent fracture toughness (K ) as a function of the loading rate. Interestingly, the data suggests a loading rate dependent size effect for the warm columnar freshwater S2 ice studied: there is a size effect at low rates, but there is no size effect at high rates. At low rates, K is higher for the large specimen (both 3m x 67 Fracture of Warm S2 Columnar Freshwater ice Under Monotonic Loading: Size and Rate Effects [PI] (a) (b) Figure 3.3. (a) Load-CMOD records for the 3m x 6m samples with loading rates as indicated in the √ √ −1 −1 figure. The 39.3kPa ms and 1.83 kPa ms data were smoothed by using moving averages (PI). (b) The pre-peak load-CMOD record. NCOD1 COD 1.5 CMOD 0.5 0 1020304050 (a) (b) −1 Figure 3.4. Load-displacement records for specimen RP4 loaded at rate 6.22 kPa ms (a) and for −1 specimen RP3 loaded at rate 57.1kPa ms (b). Both specimen had the dimensions of 0.5m x 1m. Locations, where CMOD, COD, and NCOD1 were measured, are shown in Fig. 3.1. 6m and 19.5m x 36m) than for the small specimen (0.5m x 1m), but when K is higher −1 than about 10kPa ms , the specimen size does not have an effect on K . For both the large and small specimen, K is decreasing with increasing loading rate, but this rate effect is stronger for the large specimen. This observed size effect is clear between specimen of size 0.5m x 1m and of size 3m x 6m, but there was no size effect between the 3m x 6m specimen and the largest specimen of size 19.5m x 36m. The relation between the apparent fracture toughness and the loading rate is non-linear and can be well described with a power-law relation. Fig. 3.6 shows the crack opening displacements at the crack growth initiation as a function of the loading rate. The CMOD data shows different values for each specimen size but the NCOD1 data suggests that the deformations near the crack tip are similar for the 3m x 6m and the 19.5m x 36m specimen. Both these specimen sizes appear large as compared with the small, 0.5m x 1m, specimen. This is in line with the 68 Fracture of Warm S2 Columnar Freshwater ice Under Monotonic Loading: Size and Rate Effects [PI] -1 0 1 2 10 10 10 10 Figure 3.5. Measured apparent fracture toughness as a function of loading rate. First-order power-law fits were applied separately to the data from the larger specimen (3m x 6m and 19.5m x 36m) and from the smaller specimen (0.5m x 1m) (PI). K data in Fig. 3.5. Also similarly to the K data, a rate dependent size effect was Q Q observed: at low rates, NCOD1 is larger for the large specimen, but with increasing loading rate, NCOD1 for the large specimen decreases and approaches the NCOD1 of the small specimen. For NCOD1, the large specimen display a power-law type rate effect, but the small specimens do not show a rate effect. -1 0 1 2 -1 0 1 2 10 10 10 10 10 10 10 10 (a) (b) Figure 3.6. Measured initiation crack opening displacements at the crack mouth (a) and near the crack tip (b) as a function of loading rate. First-order power-law fits were applied to the data. In (b), the power-law fit was calculated separately for the larger specimen (both 3m x 6m and 19.5m x 36m) and from the smaller specimen (0.5m x 1m) (PI). Note that a small error in PI has been fixed here: the power-law exponent in (b) for the small samples should be -0.03 instead of -0.01. 3.3.2 Nominal Strength Analysis In studies on size effect, the maximum nominal tensile strength (σ ) at the crack tip location at failure has been used as a measure for the strength (Bazant, 1984; Dempsey 69 Fracture of Warm S2 Columnar Freshwater ice Under Monotonic Loading: Size and Rate Effects [PI] et al., 1999b,a). For the edge-cracked rectangular geometry with the current loading configuration, the nominal stress is given by 2P 2 + a − 3d/2 max 0 σ = g(a ), g(a ) = (3.27) n 0 0 hL (1 − a ) where d is defined in Eq. 3.5. If a failure process was dominated by a strength criterion, such as yield, σ would be independent of the specimen size, but if a failure process was dominated by fracture, σ would decrease as a function of the size. In many quasi-brittle materials, like concrete, where a process zone in front of a crack can be identified, the failure of small specimen can be explained with a strength criterion and the failure of large specimen with a fracture criterion, as described by Bazant (1984). −0.5 For an LEFM type of fracture, σ is related to L . 1 0.2 0.8 -0.2 -0.4 0.6 -0.6 0.4 -0.8 -1 0.2 -1.2 -1 0 1 2 -1.4 10 10 10 10 -1 -0.5 0 0.5 1 1.5 (a) (b) Figure 3.7. Nominal tensile strength σ as a function of loading rate (a) and specimen size (b). First- order power-law fits were applied to the data. In (b), the data for the 0.5m x 1m and 3m x 6m specimen is obtained from the power law fits at different loading rates; the loading rate −1 of K = 0.871 kPa ms is indicated with the vertical dashed line in (a). Fig. 3.7a shows how the nominal strength (σ ) of the ice specimen is decreasing both with increasing loading rate and with increasing specimen size. The decrease of σ with rate can be described with a power-law. Then, taking data points from these power-law fits for the 0.5m x 1m and 3m x 6m specimen, Fig. 3.7b gives the relation between σ and specimen size (L) for the indicated loading rates. The decrease of σ with L can be described with a power-law where the exponent is decreasing with −0.37 −0.65 increasing rate: σ ∝ L is found for the lowest rate and σ ∝ L for the n n −0.49 highest rate. If averaging for all the data, σ ∝ L is obtained. The data in Fig. 3.7b suggests further that the rate effect is decreasing with decreasing specimen size and appears to disappear if specimen smaller than 0.5m x 1m were studied; the same result was obtained in the LEFM analysis of the data. 70 Fracture of Warm S2 Columnar Freshwater ice Under Monotonic Loading: Size and Rate Effects [PI] 3.3.3 Viscoelastic Fictitious Crack Model Analysis The viscoelastic fictitious crack model (VFCM) and the optimization procedure out- lined in Section 3.2 were used to analyze the experimental data. Fig. 3.8 shows the optimization results from RP5: convergence of the cohesive stresses (σ ,σ ,σ , and t 1 2 σ ), critical crack opening displacement (δ ), creep compliance constant (C), and 3 c objective function (F ). Similar behavior was observed for all the experiments and the results are given in Table 3.2: the number of iterations needed to reach convergence, the five control points defining the σ − δ relation (σ ,σ ,σ ,σ , δ ; Fig. 3.2b), t 1 2 3 c the crack opening displacement at crack growth initiation at X = A (δ ), the creep 0 0 compliance constant (C), the full fracture energy under FDPZ conditions (G ; Fig. 3.2b), the actual fracture energy (G ), and the process zone length (PZ; Fig. 3.2a). ac The VFCM analysis required successful measurements of CMOD, COD, and NCOD1 (Fig. 3.1) at 15 time instances during the loading. These displacements were small and challenging to measure at the initial, low load levels. Only those experiments, where the whole displacement history was measured successfully, are included in Table 3.2. Note that Table 3.1 includes data from all the 14 experiments, as the analysis in Table 3.1 is based on measured maximum values, not on the whole displacement-time record. Table 3.2. Optimization results computed using the viscoelastic fictitious crack model. ˙ 11 Experiment size K Number of σt σ1 σ2 σ3 δc δ0 C x10 G f Gac PZ σn/σt √ √ −1 2 −1 −1 (m) (kPa ms ) iterations (MPa) (MPa) (MPa) (MPa) (μm) (μm) (m N s )(N/m) (N/m) (mm) RP2 0.5x1 3.247 243 1.29 1.29 1.24 1.06 7.09 4.07 5.8 6.54 4.97 6.9 0.509 RP3 0.5x1 57.058 148 1.30 1.30 1.23 0.30 7.57 1.30 3.4 5.48 1.69 2.3 0.332 RP4 0.5x1 6.223 247 1.26 1.26 1.26 1.26 4.44 2.66 6.7 3.42 3.35 4.6 0.401 RP5 0.5x1 0.231 167 1.30 1.28 1.27 1.11 5.27 0.77 4.1 3.63 1.00 1.4 0.402 RP6 0.5x1 0.179 203 1.28 1.28 1.22 0.80 6.73 2.85 3.3 6.49 3.65 5.1 0.456 RP7 3x6 1.834 161 1.30 1.30 1.17 0.54 33.86 3.25 5.2 16.89 4.24 5.9 0.199 RP8 3x6 39.259 688 1.24 1.00 1.00 1.00 18.00 1.44 5.9 14.60 1.78 2.7 0.136 RP9 3x6 1.221 229 1.32 1.22 1.09 0.92 32.80 4.23 5.6 22.30 5.54 6.4 0.222 RP12 3x6 0.315 211 1.02 1.02 1.02 1.02 37.00 8.14 6.2 27.54 8.17 18.4 0.348 RP13 3x6 0.179 130 1.00 1.00 1.00 1.00 40.10 5.69 8.8 30.00 5.69 13.2 0.295 RP14 19.5x36 0.871 476 1.05 1.05 1.05 0.80 36.30 5.58 9.0 20.30 5.83 12.3 0.121 Figs. 3.9 and 3.10 illustrate the correspondence between the experiments and the VFCM for two experiments. The figures show the load as measured and as applied to the model, together with the response of the ice samples as measured and as obtained from the model. Both experiments show a good agreement between the model results and experimental data. This agreement supports the ability of the VFCM to describe the response of columnar freshwater S2 ice: the VFCM is able to model the non-linear 71 Fracture of Warm S2 Columnar Freshwater ice Under Monotonic Loading: Size and Rate Effects [PI] 5.4 1.25 1.2 5.2 1.15 1.1 1.05 0 1 2 3 0 2 10 10 10 10 10 10 1.8 -5 1.6 1.4 1.2 -8 0.8 -11 0 2 0 2 10 10 10 10 Figure 3.8. Optimization results of RP5. Convergence of (a) the cohesive stresses, (b) the critical crack opening displacement, (c) the creep compliance constant, and (d) the objective function after a certain number of iterations. displacement record not only at the crack mouth where the loading is applied (CMOD), but also along the crack (COD) and near the crack tip (NCOD1). The model does not account for crack propagation and thus the comparison is shown up to the peak load only. 6 300 Experiment Model 5 250 3 150 2 100 1 50 050 100 150 0 50 100 150 (a) (b) (c) Figure 3.9. Experimental and model results for RP9 (3m x 6m). (a) Load at the crack mouth, see Fig. 3.1 and Eq. (3.1). (b) Displacement - time records. (c) Load - displacement records. The CMOD data was smoothed with moving average smoothing with a span of 10% successive data points. Fig. 3.11 shows the crack opening displacement at crack growth initiation at X = A (δ ) and the actual fracture energy (G ) as functions of loading rate. Similar rate and 0 ac size effects were observed than with the linear elastic fracture mechanics analysis above. The results from the large and mid-size specimen (3m x 6m and 19.5m x 36m) are again interchangeable and higher than the results from the small specimen (0.5m x 72 Fracture of Warm S2 Columnar Freshwater ice Under Monotonic Loading: Size and Rate Effects [PI] Experiment Model 0 0 0 50 100 150 200 250 0 50 100 150 200 250 (a) (b) (c) Figure 3.10. Experimental and model results for RP14 (19.5m x 36m). (a) Load at the crack mouth, see Fig. 3.1 and Eq. (3.1). (b) Displacement - time records. (c) Load - displacement records. (The corresponding figures in PI are edited; (0,0) is added to the model plots.) 1m). This size effect is loading rate dependent and decreases with increasing rate: δ and G for the larger specimen decrease with increasing loading rate and approach ac the values obtained for the small specimen. No significant rate effect was observed for the small specimen. Power-law relations can be used to describe the decrease of δ and G with rate. ac 4 4 0 0 -1 0 1 2 -1 0 1 2 10 10 10 10 10 10 10 10 (a) (b) Figure 3.11. The crack opening displacement at crack growth initiation at X = A (a) and the actual fracture energy (b) as a function of loading rate. Viscoelastic fictitious crack model was used in the analysis. First-order power-law fits were applied separately to the data for the larger specimen (3m x 6m and 19.5m x 36m) and for the smaller specimen (0.5m x 1m) (PI). Stress-separation curves were constructed by using the five control points shown in Fig. 3.2b, and straight lines connecting the points. Fig. 3.12 shows the back-calculated σ−δ curves under FDPZ conditions and illustrates the impact of the specimen size and loading rate. With increasing separation, the σ − δ curves show initially an approximately constant stress and then extend with a softening behavior, which is getting gradually steeper. This kind of concave stress-separation curve is similar to what has been observed for sea ice by Mulmule and Dempsey (1999). The data 73 Fracture of Warm S2 Columnar Freshwater ice Under Monotonic Loading: Size and Rate Effects [PI] obtained with the larger specimen (Fig. 3.12b) illustrate that with increasing loading rate, the length of the constant part is decreasing while the slope of the softening part appears less sensitive to the loading rate. This suggest that the observed decrease in fracture energy with loading rate (Fig. 3.11) is due to a decrease in the constant part, not due to change in the softening part. At the highest rate studied, no constant part of the stress-separation can be observed, the curve shows more-or-less linear softening only. The stress-separation curves obtained with the small specimens (Fig. 3.12a) show similar behaviour, but not as clearly. The stress-separation curve includes the tensile strength (σ ) of the ice studied (Fig. 3.12); σ is given by σ at δ = 0. This value was not as strongly or clearly affected t coh v by specimen size or loading rate as the fracture parameters: for the small specimen σ ≈ 1.3 MPa, for the large specimen σ = 1.0··· 1.3 MPa, higher for the higher rates. t t The change in the shape of the σ − δ curve with loading rate is an indication of changes in the deformation and fracture processes with rate. At low rates, the approximately constant initial part of the σ − δ curve indicates that the freshwater ice deforms initially without softening and the material inside the cohesive zone is able to transfer the full tensile strength for some time with increasing load, until the deformation becomes so large that softening starts and finally leads to extension of the physical, traction-free crack. At high rates, no constant part of the curve is observed and softening initiates immediately when the critical stress is reached. This explains why freshwater ice appears brittle at high loading rates and less brittle at low rates. In fact, Fig. 3.12 represents the unattained σ − δ curves of fully-developed process zones. Although the experiments were conducted under displacement control, none of them realized a FDPZ, i.e. consumed the full fracture energy, and the experiments fractured at peak load. As illustrated in Fig. 1.5, crack growth initiation occurred when the crack was in the state shown in Fig. 1.5b. The crack did not achieve the fully-developed state shown in Fig. 1.5c; the cohesive stress (σ ) did not drop to coh zero and the crack opening displacement (δ ) did not reach the critical displacement (δ ). Only a portion of the stress-separation curve (Fig. 3.12) and the fracture energy (G , Fig. 3.2b and Table 3.1) was attained prior to the initiation of crack growth. It is possible that the very high homologous test temperatures and the inevitable grain boundary melting (Dash et al., 2006) are the reason for not attaining a FDPZ. Fig. 3.13 displays the attained σ−δ curves describing the real process zone behavior of the 0.5m x 1m (Fig. 3.13a) and the 3m x 6m and 19.5m x 36m (Fig. 3.13b) specimens at the testing conditions. These curves were deduced by interpolation from the complete stress-separation curve (Fig. 3.12) and the process zone profile deduced by the model. Generally, the specimens deformed without softening, and the process zone was transferring full tensile strength when fracture occurred. Thus, a 74 Fracture of Warm S2 Columnar Freshwater ice Under Monotonic Loading: Size and Rate Effects [PI] (a) (b) Figure 3.12. Stress-separation curves under FDPZ conditions for the 0.5m x 1m specimen (a) and for the 3m x 6m and 19.5m x 36m specimen (b) at different loading rates. The number next to each curve reflects the index of the experiment in the Legend (PI). (a) (b) Figure 3.13. The attained stress-separation curves for the (a) 0.5m x 1m specimens and for the (b) 3m x 6m and 19.5m x 36m specimens. The number next to each curve reflects the index of the experiment in the Legend [PIII]. fully developed process zone was not obtained: σ 0 and δ δ at crack growth coh 0 c initiation. The area under the σ−δ curves gives the fracture energy actually consumed, denoted as G in Table 3.2. ac Fig. 3.14a shows the upper-half crack profiles over the normalized crack length (A/L, Fig. 3.2a) for the 3m x 6m and 19.5m x 36m specimens from the experiment (defined by CMOD, COD and NCOD1, Fig. 3.1) and the VFCM. The optimized values generated by the model were relatively close to the measured values. The modelling worked better for the faster experiments (RP7 and RP8), especially at the crack mouth. However, for slower experiments (RP12, RP13 and RP14), the model underestimated the CMOD, COD and NCOD1 values. This clearly gives evidence on some effects that were more pronounced in slow experiments than fast experiments, and these effects are not accounted for in the model. The data-model misfit is mainly 75 Fracture of Warm S2 Columnar Freshwater ice Under Monotonic Loading: Size and Rate Effects [PI] due to the very high homologous test temperatures, liquidity on the gain boundaries, and the inevitable grain boundary melting (Dash et al., 2006) which have a significant influence at slow loading rates (Bell and Langdon, 1967). As discussed by Bell and Langdon (1967), under conditions of slow rate of strain, grain boundary sliding can play a very important role in the overall deformation. This concludes that the VFCM, which does not account for grain boundary sliding, works well for a range of relatively fast loading rates; slower than which results in a misfit between the measured data and the model. In addition, the model generated a better fit for the CMOD than the COD and NCOD1 values. This can be explained by the choice of the elastic modulus 1/2 in the model. In the viscoelastic characterization (Eq. 3.17, J (t) = 1/E + Ct ), E was computed as the short-time elastic modulus from the load-displacement record measured at the crack mouth. This probably led to more accuracy at the crack mouth. Another reason could be due to an increased viscoplasticity and minor viscoelasticity in the vicinity of tip. Therefore, the model worked better at the crack mouth than near the crack tip. Figs. 3.14b shows the upper-half PZ profiles by the VFCM model for the 3m x 6m specimens and the 19.5m x 36m specimen as function of the distance along the crack from the traction-free tip (X = A ). According to the VFCM, the crack-opening profile has a cusp shape with a zero gradient at X = A, with X being the position along the crack. Fig. 3.14b also gives the process zone size achieved in each experiment, as reported in Table 3.2. Fig. 3.15 shows the variation of the process zone size (PZ) with loading rate. The data from the two larger specimens (3m x 6m and 19.5m x 36m) are again interchangable and show a loading rate effect: the size of the process zone increased with decreasing loading rate. The data for the small specimens (0.5m x 1m) are scattered. Fig. 3.15 gives the same result as shown above: the size effect is loading rate dependent; the difference of the PZ values between the specimen sizes was large at low loading rates, but vanished at high loading rates. Fig. 3.16 illustrates the calculated growth of the process zone in non-dimensional form for both the small, 0.5m x 1m, specimens and the large, 3m x 6m and 19.5m x 36m, specimens. The process zone size was normalized by L and the time axis by the time to failure (t ) of each experiment (Table 3.1). The PZ - time relation is non-linear; the growth rate is increasing with time. For both the small and large specimen, PZ is increasing with decreasing rate for the high rates, but due to scatter, any trend is less clear for the lower rates. Decrease of PZ with decreasing rate appears non-physical, and thus it is possible that PZ has reached a maximum size at the lower rates; about 6mm for the smaller specimens and about 15mm for the larger specimens. 76 Fracture of Warm S2 Columnar Freshwater ice Under Monotonic Loading: Size and Rate Effects [PI] 0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 (a) 4.5 3.5 2.5 1.5 0.5 0 5 10 15 20 (b) Figure 3.14. (a) Experimental and VFCM results for the crack profiles and (b) PZ profiles by the VFCM for the 3m x 6m and 19.5m x 36m specimens. The total separation between the upper and lower crack surfaces is given by 2U ; U is half of δ which is the viscoelastic crack- v v v opening-displacement (see Fig. 1.5). X = A is the coordinate of the traction-free tip, and a is the normalized traction-free crack length (a = A /L). The number next to each curve 0 0 0 reflects the index of the experiment in the Legend [PIII]. 77 Fracture of Warm S2 Columnar Freshwater ice Under Monotonic Loading: Size and Rate Effects [PI] 1.5 0.5 -1 0 1 2 10 10 10 10 Figure 3.15. Variation of the process zone size (PZ) with loading rate and specimen size. First-order power-law fits were applied separately to the data for the larger specimen (3m x 6m and 19.5m x 36m) and for the smaller specimen (0.5m x 1m) (PI). 0.016 0.014 0.012 0.01 0.008 0.006 0.004 0.002 0 0.2 0.4 0.6 0.8 1 (a) -3 3.5 2.5 1.5 0.5 0 0.2 0.4 0.6 0.8 1 (b) Figure 3.16. The growth of the normalized process zone with normalized time for the small, 0.5m x1m, specimens (a) and for the large, 3m x 6m and 19.5m x 36m, specimens (b). The process zone size was normalized by L and the time axis by the time to failure (t ) of each experiment (Table 3.1). 78 Fracture of Warm S2 Columnar Freshwater ice Under Monotonic Loading: Size and Rate Effects [PI] 3.3.4 Notch Sensitivity Analysis Notch sensitivity is given by the ratio between the peak nominal tensile stress (σ ) at the crack tip and the tensile strength (σ ) of an uncracked specimen. This ratio is function of the brittleness number β which is used to test that a fracture test is notch sensitive and the validity of LEFM (Dempsey, 1991). Using P provided in Eq. (3.7) and σ defined in Eq. (3.27), it follows that max n = C β (3.28) n t in which K d 2π Ic β = √ , C = √ g(a ) (3.29) t n 0 a Z (d, a ) 0 1 0 σ L Note that β as used in the thesis is expressed in terms of Dempsey’s notation of K t Q (Eq. 3.7) instead of K , given that no standard exists for K . Ic Ic Table 3.2 gives the notch sensitivity values for the experiments, computed using the σ and σ values listed in Table 3.1 and Table 3.2, respectively. n t Figs. 3.17a, 3.17b, and 3.17c show the σ /σ versus the normalized crack length for n t the 0.5m x 1m, 3m x 6m, and 19.5m x 36 m specimens, respectively. Lines of constant brittleness numbers are shown. The crack length of each specimen size is indicated by the vertical dashed line. It is evident that the fracture experiments completely lose their meaning if σ = σ : a strength failure occurs prior to the attainment of a n t critical SIF. For the 0.5m x 1m and 3m x 6m specimens, that occurs for β higher 0 0 than β ≈ 0.36; for the largest specimen β ≈ 0.33 (Fig. 3.17). This is shown by the t t intersection of the horizontal dashed line of σ /σ = 1 and the vertical dashed line. n t For values of β <β , the fracture experiments are more significant for shorter cracks. However, the selection of the optimum crack length should ensure simultaneously a polycrystalline behavior and a high enough degree of brittleness (low enough σ /σ ) n t for a given specimen size. More importantly, the brittleness number can reveal the optimum specimen size so that the experiments are significant (β <β ) and suitably notch sensitive (σ /σ << 1) t n t which further restricts β . Considering size effect, previous analyses showed that the results of the two larger, 3m x 6m and 19.5m x 36m, samples were interchangeable, suggesting from Figs. 3.17b and 3.17c that a σ /σ < 0.4 is good enough to generate n t size-independent fracture results. Dempsey (1991) proposed the same ratio to ensure notch sensitivity and referred to it as the specimen shape-independent condition. It is worth noting that the 0.5m x 1m, 3m x 6m, and 19.5m x 36m plates are 77x, 462x, and 3000x the grain size, respectively. Mulmule and Dempsey (2000) speculated that sample homogeneity for first-year sea ice would be obtained if the crack-parallel specimen size harboured at least 200 d , where d is the average grain size. These av av 79 Fracture of Warm S2 Columnar Freshwater ice Under Monotonic Loading: Size and Rate Effects [PI] 1 1 0.8 0.8 0.6 0.6 0.4 0.4 0.2 0.2 0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1 (a) (b) 0.9 0.8 0.8 0.7 0.6 0.6 0.5 0.4 0.4 0.3 0.2 0.2 0.1 0 0.2 0.4 0.6 0.8 1 -1 0 1 2 10 10 10 10 (c) (d) Figure 3.17. Notch sensitivity (σ /σ ) as a function of the normalized crack length for the 0.5m x 1m n t (a), 3m x 6m (b), and 19.5m x 36 m (c) specimens. Lines of constant brittleness numbers are shown. The notch sensitivity of σ /σ = 0.4, 1 are indicated with the horizontal dashed n t lines. The normalized crack length of each specimen size is shown by the vertical dashed line. (d) Notch sensitivity (σ /σ ) as a function of the loading rate. First-order power-law n t fits were applied to the data for the 0.5m x 1m and 3m x 6m specimens (PI). experiments for very warm S2 columnar ice suggest that a crack-parallel specimen size of about 460 d is enough to achieve polycrystalline homogeneity, while 77 d av av is not. Considering the rate effect further restricts the notch sensitivity requirement. Fig. 3.17d extends this analysis by showing the variation of σ /σ as a function of loading n t rate. First-order power-law fits were applied to the data for the 0.5m x 1m and 3m x 6m specimens. The 3m x 6m specimens displayed a higher rate effect than the 0.5m x 1m specimens which satisfied the shape-independent condition only at high rates. For the 3m x 6m specimens: at low rate, the suggested σ /σ < 0.4 is valid; while at n t higher rates, a much lower ratio of σ /σ < 0.2 is necessary to ensure notch sensitivity. n t The discussion of notch sensitivity in terms of both size and rate further limits the brittleness condition to σ /σ < 0.2. A tentative conclusion is that the selection of n t the optimal specimen size and crack length, for LEFM applicability, should consider 80 Fracture of Warm S2 Columnar Freshwater ice Under Monotonic Loading: Size and Rate Effects [PI] simultaneously polycrystalline homogeneity and notch sensitivity, including the effects of size and rate. This is the first fracture study to discuss the effect of loading rate on the notch sensitivity of ice. Previous fracture studies focused only on the effect of scale on notch sensitivity (Dempsey, 1991). 3.4 Discussion The observed fracture characteristics of S2 columnar freshwater ice are reminiscent of what is expected of quasi-brittle materials, such as concrete and rock (Hillerborg, 1983). The direct consequence is that there is no validity to plane stress and plane strain idealizations, which proved very important to the development of the metals-based fracture mechanics (Broek, 1974). This implies that the thickness of the specimen (i.e. the length of the crack front) is not an important variable for the fracture behavior, as has been shown for the fracture of concrete (Hillerborg, 1983; Mindess and Nadeau, 1976). Hillerborg (1983) stated: "The microcracking of concrete is not accompanied by any substantial contraction corresponding to that which occurs when metals yield. For this reason no essential difference exists between plain strain and plane stress conditions for concrete. For concrete we thus do not need to bother about the difference between plain strain and plain stress, which causes a lot of complexity for metals. One consequence of this is that the width of the specimen (the length of the crack front) cannot be expected to be of any major importance for concrete, like it is for metals." Hillerborg’s conclusion was confirmed by experimental work done by Mindess and Nadeau (1976). Ice, being a quasi-brittle material, is expected to behave in a similar way. For long crack fronts, a triaxial state of stress does not materialize, as the tensile stress parallel to the crack front is relieved by microcracking parallel to the top and bottom surfaces or by creep, the latter taking place rapidly. The fracture experiments conducted with columnar freshwater S2 ice showed both size and rate effects which can be expressed either as a rate dependent size effect or as a size dependent rate effect. There was a size effect at the lower rates but no size effect at the higher rates. There was a rate effect for the larger specimens but a weak or no rate effect for the smaller specimens. This interrelation of size and rate effects for columnar freshwater ice is a novel observation; earlier experiments have observed rate effects (Liu and Loop, 1972; Liu and Miller, 1979; Hamza and Muggeridge, 1979; Urabe and Yoshitake, 1981; Timco and Frederking, 1986; Nixon and Schulson, 1986; Tuhkuri, 1987; Weber and Nixon, 1996a,b) or size effects (Mulmule and Dempsey, 1999; Dempsey et al., 2018, 1999b,a), but not how these two are related. In addition, the rate dependency of fracture energy has not been reported earlier: fracture energy increased with decreasing loading rate (Fig. 3.11b). Compared with the earlier studies 81 Fracture of Warm S2 Columnar Freshwater ice Under Monotonic Loading: Size and Rate Effects [PI] with freshwater S2 ice, the specimens tested here were large (L ≥ 0.5)m, very warm (≥−0.3 C), and tested in-situ. The apparent fracture toughness (K ) measured – about 100 kPa m at high rates and about 100 − 200 kPa m at low rates – is within the range measured earlier for columnar S2 freshwater ice (Dempsey, 1991). While values for K are given in Table 3.1and Fig. 3.5, the validity of K as a material parameter for ice should be discussed. K is an LEFM parameter and thus limited to be used when the material response can be idealised as linearly elastic, except in a small process zone near the crack tip, where yielding – or in the case of ice, microcracking and creep – may occur (Riedel and Rice, 1980). If such a process zone is small enough compared with the other dimensions of the specimen, fracture toughness governs crack growth; if a specimen experiences extensive yielding or creep – near the crack tip or elsewhere in the specimen – fracture toughness is not a relevant parameter. The size of the process zone in ice has been estimated with different methods. Following Riedel and Rice (1980), who studied tensile cracks in elastic-nonlinear viscous materials, a creep zone size for freshwater ice has been estimated to be in the −1 ◦ range of 0.01··· 0.5 mm at a rate of about 2 kPa ms and temperatures of −10 Cor lower (Timco and Frederking, 1986; Nixon and Schulson, 1986; Xu et al., 2004). It is important to note that the elastic power-law model of Riedel and Rice is applicable to materials where considerable ductility is present in the process zone. Another method used for ice is the method suggested by Veerman and Muller (1972), where the size of a plastic zone near a crack tip is related to the location of an apparent rotation axis of a specimen in bending. For specimen with the largest dimension of 0.4 m, a rate −1 ◦ of 2 kPa ms and temperature of −5 C, this method has suggested a process zone size of about 0.5 mm (Weber and Nixon, 1996a,b). These results of a small process zone have led to conclusions that LEFM is a suitable model for ice fracture at the rates and temperatures studied. However, the above values are an order of magnitude smaller than the about 5 mm measured here at the same rate but with larger specimens and warmer ice (Fig. 3.15). The PZ measured here for freshwater ice, about 5 − 15 mm depending on rate and specimen size, is again an order of magnitude smaller than the PZ measured for sea ice: about 25 mm for 1 m specimen and up to 150 mm for test sizes larger than 3 m (Mulmule and Dempsey, 1999). The size of the freshwater fracture process zone, as with many ice parameters, is affected by scale, rate, and temperature. Rodin and coworkers (Rodin, 2001; Wang et al., 2008) studied similar polycrystalline S2 warm ice and modelled the plate as a single grain specimen governed by the crack tip’s location and loaded through a metal-ice composite testing machine. Their model anticipated smaller PZ size than the grain size. However, the current observed PZ sizes are significant and are of the order of the grain size if not 82 Fracture of Warm S2 Columnar Freshwater ice Under Monotonic Loading: Size and Rate Effects [PI] multiple times the grain size, especially as the specimen gets larger with the 3m x 6m and 19.5m x 36 m specimens. This leads to the conclusion that the one grain ice-metal system envisaged by Rodin and coworkers is not applicable in this case. 3.5 2.5 1.5 0.5 0 102030405060 1/2 Figure 3.18. Variation of creep compliance (J = 1/E + Ct , given in Eq. 3.17) as a function of loading rate (K) for different specimen sizes (PI). (J in PI was changed to J in the thesis to distinguish between the creep compliance (J ) and the J-integral (J)). Freshwater ice is a viscoelastic material, and it can be hypothesised that the viscous deformation is less significant when the loading rate is high, and the temperature is low. Quantitively the elastic and viscous responses can be analysed by using the non-linear viscoelastic model (VFCM) and especially the creep compliance (J = 1/E + Ct ). J , calculated for each experiment by using E = E , t = t , and C from Tables 3.1 c CMOD f and 3.2, is shown in Fig. 3.18 as a function of the loading rate. Note that both E and C show some scatter but not a relation with the loading rate. As J is a function of time, a rate effect is observed, but interestingly, J appears not to have size effect–at least the size effect is much weaker than for the other parameters studied here. At high loading rates, J approaches the value of 1/E (elastic compliance), but with decreasing loading rate, the viscous component of J increases and the material response deviates −1 more and more from an elastic behavior. At loading rates K ≈ 20 kPa m s or less, J is twice or more of the elastic compliance. Thus, the warm freshwater ice studied −1 here can be considered elastic only when K is higher than about 50 kPa m s ; or the time to failure is less than 1 − 2 seconds. This strongly supports the notch sensitivity ratio (σ /σ < 0.2), suggested in Section 3.3.4, for LEFM applicability. For other n t types of ice and for other temperatures, different limit rates for elastic response should be found. Earlier studies, based on analysis of the process zone size, have concluded −1 that LEFM is a valid fracture model when K > 10 kPa m s (Urabe and Yoshitake, 83 Fracture of Warm S2 Columnar Freshwater ice Under Monotonic Loading: Size and Rate Effects [PI] 1981; Nixon and Schulson, 1986), suggesting a larger range for LEFM than suggested here. Finally, it is interesting to note that the tensile strength (σ ), obtained from the experimental data through the VFCM analysis, was about 1.2 MPa (Figs. 3.12 and 3.13) and not as strongly affected by size or rate as the fracture parameters. In earlier experiments tensile strength of ice has been observed to be a function of temperature, grain size, orientation, porosity – all constants in the experiments reported here – but also of strain rate (Schulson and Duval, 2009). This raises the question of why the tensile strength, as defined in the VFCM, does not show a clear rate effect when it has been measured before, and when the other fracture parameters do show both rate and size effects. 84 4. Fracture of S2 Columnar Freshwater Ice Under Creep/Cyclic-Recovery Loading [PII] The following chapter covers the second part of the experimental program. Load control (LC) experiments were conducted to characterize the time-dependent and fracture behavior of 3m x 6m warm and floating edge-cracked rectangular plates of warm columnar freshwater S2 ice. A program of five load control (LC) experiments was completed in the test basin (40 m square and 2.8 m deep) at Aalto University. The loading was applied in the direction normal to the columnar grains and consisted of creep/cyclic-recovery sequences followed by a monotonic ramp to fracture. The LC results were compared with the fracture results of monotonically loaded displacement control (DC) experiments of the same ice (Chapter 3), and the effect of the creep and cyclic sequences on the fracture properties were analyzed. To characterize the nonlinear displacement-load relation, Schapery’s constitutive model of nonlinear thermodynamics Schapery (1969) was applied to analyze the experimental data. A numerical optimization procedure using Nelder-Mead’s (N-M) method was implemented to evaluate the model functions by matching the displace- ment record generated by the model and measured by the experiment. The accuracy of the constitutive model is checked and validated against the experimental response at the crack mouth. This study presents the first attempt to use Schapery’s model for freshwater ice. The choice of this model for freshwater ice is motivated by the fact that the model was successfully applied to saline ice (Schapery, 1997; Adamson and Dempsey, 1998; LeClair et al., 1999, 1996) with encouraging results. The model accurately described the deformation response during load/unload applications over varying load profiles. This chapter aims to assess the time-dependent nature of warm columnar freshwater S2 ice and addresses the research questions RQ5 and RQ6 (Section 1.7). Precisely, the experiments in this study examine: 1) the extent to which the elastic, viscoelastic and viscoplastic components contribute to the ice deformation observed at the crack mouth, 2) the effects of the testing conditions on the creep stages (primary/transient 85 Fracture of S2 Columnar Freshwater Ice Under Creep/Cyclic-Recovery Loading [PII] H = 2L= 6000 Dimensions in mm NCOD3 NCOD2 L= 3000 NCOD1 COD A = 0.7L 0.5A D = 150 CMOD P P Figure 4.1. (a) Specimen geometry, edge cracked rectangular plate of length L, width H, and crack length A (PII). and steady-state/secondary) present in the ice, 3) the effects that creep and cyclic sequences have on the fracture properties; i.e. failure load and crack growth initiation displacements, and 4) the ability of Schapery’s nonlinear constitutive model to predict the experimental response. The rest of the chapter is structured as follows. In Section 4.1, a description of the applied loading profile is presented. Section 4.2 introduces Schapery’s model that is used to analyze the experiments. In section 4.3, the experimental and model results are summarized. Finally, a discussion of the results is provided in Section 4.4. The chapter has been adapted from the author’s publication (PII). 4.1 Creep/Cyclic-Recovery Fracture Experiments 4.1.1 Specimen The ice specimens tested were 3m x 6m edge-cracked rectangular plates (Fig. 4.1). 4.1.2 Creep-Recovery and Monotonic Loading Profile In two experiments, ice specimens were subjected to creep-recovery loading followed by a monotonic fracture ramp. The creep-recovery sequences consisted of four constant load applications, separated by zero load recovery periods. Each sequence was composed of alternating load/hold and release/recovery periods. Creep phases were applied at load levels of 0.4 kN, 0.8 kN, 1.2 kN, and 0.4 kN, as given by the loading signal in Fig. 4.2a. The loads were chosen low enough to avoid crack propagation and failure of the specimen. Each load-hold-unload was applied in the form of a trapezoidal wave function to avoid instantaneous load jump and drop; the 86 Fracture of S2 Columnar Freshwater Ice Under Creep/Cyclic-Recovery Loading [PII] 0 1000 2000 3000 4000 0 1000 2000 3000 4000 (a) (b) Figure 4.2. Loading consisting of (a) creep-recovery and (b) cyclic sequences followed by a monotonic fracture ramp. The number above each segment indicates the duration in s (PII). load up was applied in approximately 10 seconds and released in approximately 10 seconds. The slopes of the wave on load up and load release were 0.04kN/s, 0.08 kN/s, and 0.12 kN/s for the 0.4kN, 0.8 kN, and 1.2 kN load levels, respectively. Once at the desired hold level, the load was kept constant for a predetermined time interval. The load intervals were multiples of the hold interval for the 0.4 kN load level, Δt = 126 seconds. For the 0.8 kN and 1.2 kN load levels, the time interval was doubled and quadrupled: 2Δt = 252 seconds and 4Δt = 504 seconds, respectively. 1 1 The four zero load recovery periods, separating the creep load periods, were also function of Δt . Three recovery periods were held at zero load level for 5Δt = 630 1 1 seconds, but the last recovery period was maintained for a longer interval of 10Δt = 1260 seconds. Immediately following the creep and recovery loading sequences, the specimen was loaded monotonically to failure on a load-controlled linear ramp. The ramp up to the peak load and unloading were each applied over an interval of Δt . 4.1.3 Cyclic-Recovery and Monotonic Loading Profile In three experiments, ice specimens were loaded with cyclic-recovery sequences followed by a fracture ramp, as shown in Fig. 4.2b. The cyclic-recovery loading consisted of 3 sequences, each being composed of four fluctuating loads, at the levels of 0.4 kN, 0.8kN, and 1.2 kN. Each cyclic sequence continued for a constant time interval Δt = 480 seconds. The slopes of the wave on the load up and load release were 1/150 kN/s, 1/75 kN/s, and 1/50 kN/s for the 0.4kN, 0.8kN, and 1.2 kN load levels, respectively. The 0.4kN, 0.8kN, and 1.2 kN cyclic load periods were followed by zero load recovery periods of 1.25Δt = 600 seconds, 1.25Δt = 600 seconds, and 2 2 2.5Δt = 1200 seconds, respectively. 87 Fracture of S2 Columnar Freshwater Ice Under Creep/Cyclic-Recovery Loading [PII] At the completion of the cyclic-recovery loading sequences, the specimen was loaded to failure by a monotonic linear ramp. The ramp up to the peak load and unloading were each applied over an interval of 0.25Δt = 120 seconds. 4.2 Nonlinear Time-Dependent Modeling of S2 Columnar Freshwater Ice The model applied in this section to characterize the nonlinear elastic/ viscoelastic/ viscoplastic response of S2 columnar freshwater ice was presented by Schapery; it was used to model the time-dependent mechanical response of polymers in the nonlinear range under uniaxial stress-strain histories (Schapery, 1969). Schapery’s stress-strain constitutive equations are derived from nonlinear thermodynamic principles, and are very similar to the Boltzmann superposition integral form of linear theory (Flügge, 1975). Schapery’s model represents the material as a system of an arbitrarily large number of nonlinear springs and dashpots. The equations in this section are presented in terms of load and displacement instead of the original stress-strain relations. The notations of the original equations in (Schapery, 1969) are modified to bring out similarity between all the equations in the paper. When the applied loads are low enough, the material response is linear. For an arbitrary load input, P = P(t) applied at t = 0, Boltzmann’s law approximates the load by a sum of a series of constant load inputs and describes the linear viscoelastic displacement response of the material using the hereditary integral in a single integral constitutive equation. The Boltzmann superposition principle states that the sum of the displacement outputs resulting from each load step is the same as the displacement output resulting from the whole load input. If the number of steps tends to infinity, the total displacement is given as: dP δ(t) = C P + ΔC(t − τ) dτ, (4.1) dτ where C is the initial, time-independent compliance component and ΔC(t) is the transient, time-dependent component of compliance. Turning now to nonlinear viscoelastic response, Schapery developed a simple single-integral constitutive equation from nonlinear thermodynamic theory, with either stresses or strains entering as independent variables (Schapery, 1969). Using load as the independent variable, the displacement response under isothermal and uniaxial loading takes the following form: d(g P) δ(t) = g C P + g ΔC(ψ − ψ ) dτ, (4.2) 0 0 1 dτ 88 Fracture of S2 Columnar Freshwater Ice Under Creep/Cyclic-Recovery Loading [PII] where C and ΔC are the previously defined components of Boltzmann principle, ψ and ψ are the so-called reduced times defined by: t τ dt dt ψ = and ψ = ψ(τ) = (4.3) a [P(t )] a [P(t )] 0 P 0 P and g , g , g , and a are nonlinear functions of the load. Each of these functions 0 1 2 P represents a different nonlinear influence on the compliance: g models the elastic response, g the transient response. g the loading rate, and a is a time scale shift 1 2 P factor. These load-dependent properties have a thermodynamic origin. Changes in g , g , and g reflect third and higher order stress-dependence of the Gibbs free energy, 0 1 2 and changes in a are due to similar dependence of both entropy production and the free energy. These functions can also be interpreted as modulus and viscosity factors in a mechanical model representation. In the linear viscoelastic case, g = g = g = 0 1 2 a = 1, and Schapery’s constitutive equation (4.2) reduces to Boltzmann’s equation (4.1). Equation (4.2) contains one time-dependent compliance property, from linear vis- coelasticity theory, ΔC and four nonlinear load-dependent functions g , g , g , and a , 0 1 2 P which reflect the deviation from the linear viscoelastic response, that need to be eval- uated. Schapery’s model uses experimental data to evaluate the material property functions in Eq. (4.2). Lou and Schapery (1971) outlined a combined graphical and numerical procedure to evaluate these functions. In their work, a data-reduction method was applied to evaluate the properties from the creep and recovery data. Pa- panicolaou et al. (1999) proposed a method capable of analytically evaluating the material functions using only limiting values of the creep-recovery test. Numerical methods are also employed and are the most commonly used techniques; they are based on fitting the experimental data to the constitutive equation (LeClair et al., 1999). In the current study, a numerical-experimental procedure is adopted. An optimization procedure is applied using the Nelder-Mead (N-M) method (Nelder and Mead, 1965) to back-calculate the values that achieve the best fit between the model and the experi- mental data. To avoid multiple fitting treatments of data and account for the mutual dependence of the functions, the properties were determined from the full data. This avoided errors that may result from separating the data into parts and estimating the functions independently from different parts. This approach of fitting a model with experimental data is common in fracture models with several parameters. Schapery (1997) later updated his formulation. He added a viscoplastic term to account for the viscoplastic response of the material and stated that the total compli- ance can be represented as the summation of elastic, viscoelastic, and viscoplastic components. Adamson and Dempsey (1998) applied Schapery’s updated constitutive equation to model the crack mouth opening displacement of saline ice in an experi- mental setup similar to the current study. The theory represents the displacement at the 89 Fracture of S2 Columnar Freshwater Ice Under Creep/Cyclic-Recovery Loading [PII] crack mouth (δ ) as the sum of elastic, viscoelastic, and viscoplastic components: CMOD e ve vp δ = δ + δ + δ (4.4) CMOD CMOD CMOD CMOD where δ = g C P (4.5) 0 e CMOD d g P ve δ = g C (ψ − ψ ) dτ (4.6) 1 ve CMOD dτ vp δ = C g Pdτ (4.7) vp 3 CMOD In the above equations, ψ and ψ are defined in (4.3). g , g , g , g , and a are nonlinear 0 1 2 3 P load functions to be determined. The coefficients C , C , and C are the elastic, e ve vp viscoelastic, and viscoplastic compliances, respectively. Schapery’s equation has been developed for uniaxial loading. The response of the test specimen is dominated by the normal stresses at the direction normal to the X-axis, ahead of the crack (Fig. 4.1). This stress state can be approximated as uniaxial in the same way as in beam bending; the stress is uniaxial tension at the crack tip and then changes linearly. Thus, Schapery’s equations are used to analyze the experimental data. Few assumptions are applied at this point and are based on the choices made by Adamson and Dempsey (1998). For ice, the elastic displacement is linear with load; this immediately leads to g = 1. Schapery stated that g = a = 1 if the instantaneous jump and drop in the 0 1 P displacement are equal (Schapery, 1969). Examination of the current data shows that this condition is not valid, and the functions need to be evaluated. Accordingly, the following approximations are employed: a b−1 c−1 d g ∝ P ; g ∝ P ; g ∝ P ; a ∝ P (4.8) 1 2 3 P From Eq. (4.3): dt ψ − ψ = (4.9) a [P(t )] The viscoelastic compliance is assumed to follow a power law in time with a fractional exponent n. This gives: C (β) ≈ κβ (4.10) ve Incorporating each of these conditions, the total displacement is expressed as t t b t dt d[P(τ)] a c δ = C P + κP dτ + C P dτ (4.11) CMOD e vp [P(t )] dτ 0 τ 0 where δ , P and t are in m, N, and seconds, respectively. It follows from Eq. CMOD 4.11 that two unknown parameters (C , and C ), one unknown constant (κ), and five e vp unknown exponents (a, b, c, d, and n) need to be determined. As previously mentioned, 90 Fracture of S2 Columnar Freshwater Ice Under Creep/Cyclic-Recovery Loading [PII] the problem is optimized through the N-M technique, by minimizing the objective function F given by the difference between the model and data, as shown in Eq. 4.12. The components of the total displacement were computed and optimized using MATLAB. A positive constraint was applied to the model variables. Initial guesses of the exponents on the load and time functions were assumed based on previous work on saline ice. The optimized values were then obtained by comparing the model response and the experimentally measured response over the full length of the test up to crack growth initiation. F = arg min M (C , C ,κ, a, b, c, d, n) − D (4.12) i e vp i C ,C ,a,b,... e vp i=1 where M and D refer to the CMOD values given by the model (Eq. 4.11) and the i i experimental data, respectively. . is the Euclidean norm of a vector. N is the number of data points (≈ 2e6 points). This problem is typically called a least-squares problem when using the Euclidean norm. It is a convex problem because F is a convex function and the feasible set is convex. Thus, the optimization algorithm will converge to the global optimal solution. As mentioned earlier, Schapery’s model originated from the thermodynamic theory. The model is not physically-based, and its parameters are not linked to the microstruc- tural properties of the ice (dislocation density, grain size, ...). In addition, the analysis does not account for the formation of fracture process zone in the vicinity of the crack tip. Schapery’s formulation models the experimental response until crack growth initiation and does not account for crack propagation. Table 4.1. Measured experimental data and computed results for the LC experiments. ˙ ˙ Experiment Type LH A hE E E E E P t CMOD CMOD NCOD1 NCOD1 0 1 2 3 4 f max f −1 −1 (m) (m) (m) (mm) (GPa) (GPa) (GPa) (GPa) (GPa) (kN) (s) (μm) (μms )(μm) (μms ) RP15 creep 3 6 2.1 364 6.6 6.7 7.3 7.4 6.9 5.8 68.2 320.1 4.7 53.6 0.8 RP16 creep 3 6 2.1 385 5.6 5.8 7.6 - 6.0 3.8 42.8 228.2 5.3 49.1 1.1 RP17 cyclic 3 6 2.1 407 6.5 - 7.6 - 6.6 4.5 49.3 173.7 3.5 30.0 0.6 RP18 cyclic 3 6 2.1 408 - - - - 5.3 3.9 40.1 143.7 3.6 28.5 0.7 RP19 cyclic 3 6 2.1 412 - 7.0 6.6 - 6.3 6.3 52.5 221.4 4.2 44.0 0.8 91 Fracture of S2 Columnar Freshwater Ice Under Creep/Cyclic-Recovery Loading [PII] 4.3 Results 4.3.1 Experimental and Modelling Results This section presents the results measured and computed for the LC experiments. The current results are compared with the fracture results of monotonically loaded DC experiments of the same ice and same specimen size (3m x 6m, Chapter 3). The main aim is to elucidate the effect of creep and cyclic sequences on the fracture properties. 4.3.2 Effect of the Creep and Cyclic Sequences on the Fracture Properties Table 4.1 shows the measured and computed parameters for the LC experiments. P is the measured peak load which is also the failure load. t represents the time max f to failure, computed from the fracture ramp. CMOD is measured at crack growth initiation. CMOD indicates the displacement rate at the crack mouth and is obtained by dividing CMOD by the failure time. Similarly, NCOD1 (see Fig. 4.1) represents the displacement at crack growth initiation near the initially sharpened crack tip. NCOD1 indicates the displacement rate in the vicinity of the tip and is obtained by dividing NCOD1 by the failure time. Fig. 4.3 gives the results of the peak load P , crack mouth opening displacement max CMOD, and near crack-tip opening displacement NCOD1 as a function of the loading time for the DC experiments (Chapter 3) and the current LC experiments. In these subplots, first-order power-law fits were applied to the data of the DC experiments. The LC values lie above, below, and along the DC fit. No clear effect of creep and cyclic loading on the fracture properties was detected. Figs. 4.4a and 4.4b show the experimental load versus the crack opening displace- ment at the crack mouth for the DC and the LC experiments, respectively. Fig. 4.4c displays a zoomed view of the fracture ramp of the LC experiments. Comparing the failure loads of the DC and LC experiments indicates that the failure loads, of experi- ments with comparable loading rates, were similar. Therefore, in these experiments, the creep and cyclic sequences had no influence on the failure load. Table 4.1 presents several elastic moduli for each experiment. The elastic moduli were calculated from the load-CMOD record usin Eq. 3.16 (Chapter 3, Section 3.1). For the creep experiments (RP15 and RP16), this procedure is repeated for the four creep cycles, resulting in E , E , E , E , and for the fracture ramp, resulting in E . 1 2 3 4 f Similarly for the cyclic experiments (RP17, RP18, and RP19), the moduli calculation was done for the last cycle of each cyclic sequence, giving steady state moduli E , 92 Fracture of S2 Columnar Freshwater Ice Under Creep/Cyclic-Recovery Loading [PII] 1 50 0 1 2 3 0 1 2 3 10 10 10 10 10 10 10 10 (a) (b) 0 1 2 3 10 10 10 10 (c) Figure 4.3. Experimental results for the (a) peak load P , (b) crack mouth opening displacement max CMOD and (c) near crack tip opening displacement NCOD1 at crack growth initiation, as a function of time to failure t for the monotonically-loaded DC experiments (Chapter 3) and the creep/cyclic and monotonically-loaded LC experiments (PII). (a) (b) (c) Figure 4.4. Measured load versus CMOD for the (a) DC experiments (Chapter 3), (b) LC experiments, and (c) LC experiments up to the peak load (PII) 93 Fracture of S2 Columnar Freshwater Ice Under Creep/Cyclic-Recovery Loading [PII] E , E , and for the fracture ramp, resulting in E . Some of the values are missing, 2 3 f caused by the fact that the initial portion of the associated load-CMOD curve was very noisy. The values of the elastic moduli calculation for the creep/cyclic sequences and fracture ramps were fairly constant upon load application, as shown by the linear load-delection curves in Figs. 4.4c, 4.5a, and 4.5b. This linearity justifies the choice of g = 1 in the elastic CMOD component in Eq. 4.5. Table 3.1 in Chapter 3 presents the elastic modulus (E ) calculated at the crack CMOD mouth for the DC experiments; E is similar to E in Table 4.1 here; both values CMOD f lie within the same range. Therefore, the creep and cyclic sequences preceding the fracture ramp did not affect the load-CMOD prepeak behavior. However, the sequences affected the post-peak response as can be distinguished from Fig. 4.4b which displays a longer decay behavior than Fig. 4.4a. The gradual decay of the load portrays the time dependency in the behavior of freshwater ice. Viscoelastic response Viscoplastic response ௩ CMOD େୈ (a) (b) (c) Figure 4.5. Load versus CMOD over the (a) creep-recovery cycles for RP15 and the (b) cyclic-recovery sequences for RP17. (c) Schematic illustration of the hysteresis load-displacement diagram. The whole of the hysteresis loop area is the energy loss per cycle. The dashed area is the part of that total that is due to the viscoelastic mechanism and the rest is due to viscous processes (PII). 4.3.3 Ice Response Under the Testing Conditions Fig. 4.6 shows the experimental results for RP16: the applied load and the crack opening displacements at the crack mouth (CMOD), halfway along the crack (COD), and 10 cm behind the tip (NCOD1) (see Fig. 4.1). Similarly, Fig. 4.7 shows the experimental response for RP17. The time-dependent nature of of the ice response is evident. A complete load-CMOD curve was obtained during loading and unloading for each experiment of Table 4.1, indicating stable crack growth. It is clear from Figs. 4.6b and 4.7b that the CMOD, COD, and NCOD1 displace- ments were composed mainly of elastic and viscoplastic components. No significant viscoelasticity was detected in the displacement-time records for all the experiments. The primary (transient) creep stage was almost absent or instantaneous. The load Load Fracture of S2 Columnar Freshwater Ice Under Creep/Cyclic-Recovery Loading [PII] 4 250 0 0 0 1000 2000 3000 4000 0 1000 2000 3000 4000 (a) (b) (c) 10.3 10.4 10.5 10.6 Time (d) (e) Figure 4.6. Experimental results for RP16. (a) Load at the crack mouth, see Fig. 4.1. (b) Displacement - time records. (c) Load - displacement record. (d) Typical response of a Maxwell model, consisting of a nonlinear spring and nonlinear dashpot, to a constant load step. (e) CMOD vs time plot showing the first creep sequence. [PII]. sequences were characterized by a non-decreasing displacement rate at all levels. The displacement-time slope was linear and constant, indicating that the secondary/steady- state creep regime dominated during each load application. Although the recovery time was longer than the loading time, ≥ 1.25Δt (Creep experiment, Fig. 4.2a and Section 4.1.2) and ≥ 1.25Δt (Cyclic experiment, Fig. 4.2b and Section 4.1.3), the recovery (unload) phases consisted mainly of an elastic recovery (instantaneous drop) and unrecovered viscoplastic displacement. The behavior as observed resembles the response of a Maxwell model composed of a series combination of a nonlinear spring and nonlinear dashpot (Fig. 4.6d). There is no delayed elastic recovery, but there is the elastic response and a permanent deformation. It is important to note that the vertical-looking jumps in the creep and cyclic sequences (Figs. 4.6b and 4.7b) are real but not vertical. It is only a scaling issue; the scale makes them appear as instantaneous with zero displacement. However, zooming in (Fig. 4.6e) reveals that the jumps are just steep, reflecting the immediate response to load. Figs. 4.5a and 4.5b support the same analysis. Unlike the viscoelastic response (Fig. 4.5c) which displays no residual displacement in the loading and unloading hysteresis diagram, the current load-CMOD plots showed large permanent displacement after each loading cycle. This concludes that the response of columnar freshwater S2 ice in these experiments was overall elastic-viscoplastic. Displacement Fracture of S2 Columnar Freshwater Ice Under Creep/Cyclic-Recovery Loading [PII] 0 1000 2000 3000 4000 0 1000 2000 3000 4000 (a) (b) (c) Figure 4.7. Experimental results for RP17. (a) Load at the crack mouth, see Fig. 4.1. (b) Displacement - time records. (c) Load - displacement record (PII). 4.3.4 Nonlinear Modelling Analysis The nonlinear theory, outlined in Section 4.2, was used to analyze the experiments. Modelling the visoelastic term (second term of Eq. 4.11) proved to be very challenging. Instead, a simplified version was modelled by setting a = g = 1. The results of the p 2 initial optimization trials confirmed the previous analysis; the viscoelastic component ve δ had no effect on the final fit between the data and the model. The optimization CMOD −18 algorithm fine-tunedκ (Eq. 4.11) to a very small number (10 ), indicating that the best model-data fit is attained when the viscoelastic term goes to zero (Fig. 4.8b). The final optimization runs were carried out by considering the elastic and viscoplas- tic components (first and last terms of Eq. (4.11)) only. This resulted in 2 parameters, C and C , and one exponent c, that need to be optimized. The optimization con- e vp verged results are given in Table 4.2: C , C and c. For all the experiments, the e vp %reduction of the objective function exceeded 95% and about 110 iterations were needed to reach convergence. A value of c = 1 for the viscoplastic load function provided the best fit between the model and the experiment at all load levels over the total experimental time up to the peak load. The final compliance values of the −8 −1 elastic and viscoplastic components were in the ranges 1.8-3.8 x10 mN and 0.2-1 −10 −1 −1 x10 mN s . respectively. Table 4.2. Optimization results computed using Schapery’s model. 8 10 experiment t C x10 C x10 c f e vp −1 −1 −1 (s) (mN ) (mN s ) RP15 68.2 3.330 1.061 1 RP16 42.8 3.845 0.974 1 RP17 49.3 2.637 0.512 1 RP18 40.1 1.861 0.209 1 RP19 52.5 2.775 0.938 1 96 Fracture of S2 Columnar Freshwater Ice Under Creep/Cyclic-Recovery Loading [PII] Figs. 4.8 and 4.9 give the model results, obtained using Eqs. (4.4-4.11), and the experimental results for experiments RP16 and RP17, respectively. Figs. 4.8a and 4.8b show the measured load and the load applied to the model and the measured CMOD-time record compared to the response of the model, respectively for RP16. Figure 4.9 shows similar plots for experiment RP17. experiment RP17 showed a rather good model-experiment fit for the three cyclic-recovery sequences over the load and unload periods. The model succeeded to follow the increasing and decreasing load levels and the corresponding recovery phases. The experimental response for the creep-recovery experiment RP16 appeared to generally conform to the model results, but the model overestimated the recovery displacement in the first two cycles. It is unclear to the authors why the model did a better job in fitting the cyclic-recovery than the creep-recovery sequences. This probably hints at some mechanisms that took place in the creep-recovery experiments and are not accounted for by Schapery’s model. Schapery’s model has been tested for creep-recovery sequences of saline ice with an increasing load profile (Schapery, 1997; Adamson and Dempsey, 1998; LeClair et al., 1999, 1996). This is the first application of the model with a load profile of increasing and decreasing load levels (Fig. 4.2). Experiment Model with viscoelasticity Model with no viscoelasticity 0 1000 2000 3000 4000 0 1000 2000 3000 4000 5000 (a) (b) Figure 4.8. Experimental and model results for RP16. (a) Load at the crack mouth, see Fig. 4.1. (b) CMOD - time records: Schapery’s model with and without viscoelasticity vs the experimental data (PII). Considering the fracture ramp, Schapery’s nonlinear equation succeeded to model the monotonic displacement response up to crack growth initiation perfectly well for all the experiments. As previously mentioned, the model does not account for crack propagation, so modeling was applied until the peak load. The model was also successful in predicting the critical crack opening displacement values at the failure load. Thus, the model gives a very close prediction of the experimental data over the whole loading profile up to the failure load. The other experiments displayed the same experiment-model agreement. In this study, Schapery’s constitutive model is tested for the first time for freshwater 97 Fracture of S2 Columnar Freshwater Ice Under Creep/Cyclic-Recovery Loading [PII] Experiment 160 Model 0 1000 2000 3000 4000 0 1000 2000 3000 4000 (a) (b) Figure 4.9. Experimental and model results for RP17. (a) Load at the crack mouth, see Fig. 4.1. (b) CMOD - time records (PII). ice. The match between the model and the measured data, especially for the cyclic- recovery experiments, provides a firm support of the ability of Schapery’s constitutive model to describe the time-dependent response of columnar freshwater S2 ice up to crack growth initiation. Figs. 4.10a and 4.10b show the contribution of each individual model component, elastic and viscoplastic, to the total CMOD displacement, for RP16 and RP17, respectively. As mentioned earlier, the elastic and viscoplastic components account for the total deformation. For RP16, the viscoplastic component dominated over the elastic component. For RP17, the elastic and viscoplastic components contributed equally to the total displacement. The applicability of the proposed model and the fitted parameters are limited to the studied ice type, geometry, specimen size, ice temperature, and the current testing conditions. Variation in the operating conditions will change the dominant deformation mechanisms and the ice behavior; and accordingly, new model parameters would be needed to adapt to the new response. 100 80 0 0 0 1000 2000 3000 4000 0 1000 2000 3000 4000 (a) (b) Figure 4.10. Contribution of each individual model component to the total CMOD displacement for (a) RP16 and (b) RP17 (PII). 98 Fracture of S2 Columnar Freshwater Ice Under Creep/Cyclic-Recovery Loading [PII] 4.4 Discussion Interestingly, the ice behavior in the current study differs from previous experimen- tal creep and cyclic work on freshwater ice. Large delayed elastic or recoverable component has been previously observed. Several researchers performed creep ex- periments on granular freshwater ice at lower temperatures (Mellor and Cole, 1981, 1982, 1983; Cole, 1990; Duval et al., 1991) and reported considerable recovery. Duval conducted torsion creep experiments on glacier ice at a similar testing temperature of -1.5 C (Duval, 1978). When unloaded, the ice exhibited creep recovery. According to his analysis: during loading, the internal stresses opposing the dislocation motion increases; upon unloading, the movement of dislocations produced the reversible defor- mation and is caused by the relaxation of internal stresses. Sinha (1978, 1979) studied columnar-grained freshwater ice and concluded that the high-temperature creep is associated with grain boundary sliding. Cole (1995) developed a physically-based constitutive model in terms of dislocation mechanics and quantified two mechanisms of anelasticity: dislocation and grain boundary relaxations. He demonstrated that the increased temperature sensitivity of the creep properties of ice within a few degrees of the melting point is due to a thermally induced increase in the dislocation density (Cole, 2020). The question then arises as to why warm columnar freshwater ice tested here showed no significant delayed elastic effect, and the microstructural changes were mainly irreversible upon unloading? The measured ice response is a novel result for any type of ice. It is important to emphasize that in comparison with earlier freshwater ice studies, the tested samples are very warm and large. Viscoelasticity normally happens due to the elastically- accommodated grain boundary sliding. Upon loading, internal stresses build up at local stress concentrations in the grain boundary geometry (triple points and grain boundary ledges). Assuming there is no microcracking, the growing stress impedes further grain boundary sliding and causes sliding in the reverse direction, giving rise to the recoverable component after unloading. However, in the present case, the measurements showed large unrecoverable deformation. Several reasons can be discussed, related to the ice temperature, microstructure, and nonlinear mechanisms in the process zone. Concerning the effect of temperature: the warmer the temperature, the more liquid on the grain boundary. The high homologous test temperature (top ice surface ≈ -0.3 C) causes liquidity on the gain boundaries (Dash et al., 2006). The intergranular melt phase on the grain boundary renders the ice as two-phase polycrystal and significantly influences the creep and recovery response. In fact, the grain boundary sliding then consists theoretically of two processes: 1) the sliding of grains over one another 99 Fracture of S2 Columnar Freshwater Ice Under Creep/Cyclic-Recovery Loading [PII] and 2) the squeezing-in/out of the liquid between adjacent grains (Muto and Sakai, 1998). The shear behavior of the liquid film is function of its properties (thickness and amount). The presence of this liquid at the triple points and the boundary acted as a resisting obstacle for the grains to shear and deform back to their original form, causing the viscoplastic deformation. The microstructure (grain size, crystalline texture) could be another contributing factor. Sinha (1979) developed a nonlinear viscoelastic model, incorporating the grain size effect, to describe the high-temperature creep of polycrystalline materials. He concluded that delayed elastic strain exhibits an inverse proportionality with grain size. This suggests that the grain size (3-10 mm, Fig. 2.5b) of the ice samples is coarse enough not to produce any measurable viscoelastic deformation under the testing conditions. It is also probable that for this grain size, there was not enough local concentration points to arrest the grain boundary sliding and drive the recoverable and reverse sliding. In addition, Gasdaska (1994) discussed that regularly ordered and packed microstructures limit the amount of sliding and rearrangement and lead to less anelastic strain. The ice growth in the Aalto Ice tank was very controlled and resulted in homogeneous ice sheet. Knauss presented a thorough review of the time-dependent fracture models available to date (Knauss, 2015). The essence of the models is based on modelling the behavior in a finite cohesive/process zone which is attached to the traction-free crack tip. The one-parameter fracture mechanics encompassed by the apparent fracture toughness is not applicable (Dempsey et al., 2018). It is believed that the mechanisms taking place in the process zone play an influencing role in the current experiments. The nonlinearity in the fracture zone relieved the internal stresses that would ordinarily accommodate the grain boundary sliding and generate some viscoelastic deformation upon unloading. Thus, any microstructural damage that occurred during loading manifested as permanent deformation at the end of the experiment. It is noteworthy that the earlier studies used test sizes which are smaller than the plate size used here. It was shown in the DC fracture experiments (Chapter 3) that scale had an effect at the tested loading rates. It is probable that the specimen size influenced the time-dependent deformation of freshwater ice. The current experiments suggest that for the large sample size and the kind of ice studied (very warm freshwater ice) under the loading applied, viscoelasticity is not an important deformation component. The experimental results support this prediction, but more experiments are needed to make more general conclusions. All the above-mentioned factors might have contributed to the measured elastic- viscoplastic response. However, the question as to which factor influenced mostly the behavior is an important research question that requires more testing programs. 100 Fracture of S2 Columnar Freshwater Ice Under Creep/Cyclic-Recovery Loading [PII] Testing the effect of each factor separately requires a set of experiments that considers this factor while keeping all the other conditions fixed. 101 Fracture of S2 Columnar Freshwater Ice Under Creep/Cyclic-Recovery Loading [PII] 102 5. Fracture Energy of Columnar Freshwater Ice: Influence of Loading Type, Loading Rate and Size [PIII] This chapter addresses the research question RQ7 (Section 1.7). It investigates the influence of loading type, loading rate, and specimen size on the fracture energy of columnar freshwater S2 ice for the DC (Chapter 3) and LC (Chapter 4) experiments. Different methods for computing the fracture energy were applied and compared. The apparent fracture energy at crack growth initiation was obtained via Rice’s J-integral expression (J ) modified to be applicable to the special case of a deeply cracked edge-cracked plate as well as via a viscoelastic fictitious crack analysis (G ). The VFCM work-of-fracture (W ) was also evaluated. Both J and W were measured from f Q f the load-displacement record at the crack mouth. G was obtained from the VFCM back-calculated stress-separation (σ − δ) relation within the fracture process zone. The rest of the chapter is structured as follows. Section 5.1 introduces the fracture energies and explains how they are obtained. In section 5.2, the computed results from the DC and LC experiments are summarized and analyzed. The chapter has been adapted from the author’s publication (PIII). 5.1 Fracture Energies 5.1.1 Fracture Energy at Crack Growth Initiation via the J-integral (J ) The J-integral, formulated by Rice (1968), approximates the crack-tip stress-strain fields in a linear or nonlinear elastic material, and thus becomes more difficult to interpret once unloading occurs in the specimen. Consequently, the J-integral is used for specimens at the initiation of crack growth (Mindess et al., 1977; Homeny et al., 1980). Therefore, Rice’s J-integral gives the apparent fracture energy at crack growth initiation (J ), i.e. the energy required to initiate fracture. Later, Rice et al. (1973) derived the J-integral expression for a deeply-notched edge-cracked plate. His approximation is essentially exact if the fabricated notch is 103 Fracture Energy of Columnar Freshwater Ice: Influence of Loading Type, Loading Rate and Size [PIII] deep enough, with just a short ligament subjected to bending: crack J = Pdδ (5.1) Q crack h(L − A ) 0 0 where h is the plate thickness, L − A the remaining uncracked ligament (see Fig. 3.1), P the applied load, and δ is the crack-mouth-opening-displacement (CMOD). crack The integral is simply the area (A ) under the load-CMOD curve up to crack growth initiation. It represents the work done in loading; where the elastic deformations with no crack present should be eliminated from the calculations in Eq. 5.1. J has been used extensively in metals (Begley and Landes, 1972; Landes and Begley, 1972; Bucci et al., 1972; Landes and Begley, 1974; Clarke et al., 1976; Logsdon, 1976; Yoder and Griffis, 1976; Griffis and Yoder, 1976; Homeny et al., 1980), concrete (Mindess et al., 1977), and ceramics (Sakai et al., 1988). Apparently, few attempts have been made to estimate this quantity for any ice type. To my knowledge, the first effort for freshwater ice was reported by Li et al. (1996). They used the photoelastic method and the speckle technique to measure the stress and displacement fields necessary to obtain the J-integral. Unfortunately, they could not test for the critical J because of limitations of the testing device. Eq. 5.1 is a good approximation for the current experiments because the initially- sharpened crack length is sufficiently long: A ≈ 70%L (0.5m x 1m and 3m x 6m plates) and ≈ 75%L (19.5m x 36m plate). The elastic displacements of the load points of the uncracked specimen are negligible compared to the displacements due to the crack (Rice et al., 1973). Therefore, the total raw area under the load-CMOD curve up to the maximum load is considered, as portrayed in Fig. 5.1a. 5.1.2 Work-of-Fracture (W ) The work-of-fracture (W ), as used by Nakayama and others (Nakayama, 1964; Tattersall and Tappin, 1966; Homeny et al., 1980; Sakai et al., 1988; Bažant, 1996; Dempsey et al., 2012), represents the fracture energy required to propagate a crack completely through a specimen. It is calculated from the total work done in splitting the plate, as determined from the area under the entire load-CMOD record (A , see Fig. 5.1b), i.e. W = (5.2) h(L − A ) It is worth noting that J and W provide a more accurate analysis than the apparent Q f fracture toughness (K , Section 3.1 and Eq. 3.7). The K value depends only on Q Q the peak load; however, evaluation of J and W depends on the peak load and the Q f shape of the load-CMOD curve. The ratio of the work-of-fracture to the apparent 104 Fracture Energy of Columnar Freshwater Ice: Influence of Loading Type, Loading Rate and Size [PIII] (a) (b) (c) (d) Figure 5.1. Description of the area used for (a) J (A ) and (b) W (A ) calculations. Load versus Q J f W Q f crack mouth opening displacement for (c) RP4 and (d) RP12, illustrating the A and A J W Q f areas [PIII]. Figure 5.2. Description of the area used for J (A ) calculations for metals, as defined by the ASTM Q J E1820-20b. Adapted from ASTM E1820-20b AST (2020). Fracture Energy of Columnar Freshwater Ice: Influence of Loading Type, Loading Rate and Size [PIII] fracture energy at crack growth initiation (W /J ) gives information about the energy f Q absorbed during crack extension. 5.1.3 Fracture Energy at Crack Growth Initiation via the VFCM (G ) VFCM G represents the fracture energy at crack growth initiation obtained via a vis- VFCM coelastic cohesive analysis. A viscoelastic fictitious crack model (VFCM) was applied to model the experimental data of the monotonic experiments (Chapter 3). G is VFCM measured from the area under the σ − δ curve (Fig. 3.13) attained in each experiment. G corresponds to G in Table 3.2 of Chapter 3. VFCM ac 5.2 Results and Discussion Table 1 gives the dimensions of the ice samples together with the measured (L, H, A , h, P , t ) and computed (G , J , W , W /J ) results. max f VFCM Q f f Q Table 5.1. Specimen dimensions, measured data, and results computed for the DC and LC experiments. Experiment Type LH A hP t G J W W /J 0 max f VFCM Q f f Q m m m mm kN s N/mN/mN/m RP1 DC 0.5 1 0.35 350 1.7 55 - 1.2 3.3 2.6 RP2 DC 0.5 1 0.35 350 2.3 52 5.0 2.4 3.7 1.5 RP3 DC 0.5 1 0.35 371 1.6 2 1.7 1.4 3.3 2.3 RP4 DC 0.5 1 0.35 376 1.9 21 3.4 1.3 3.9 2.9 RP5 DC 0.5 1 0.35 402 2.1 572 1.0 1.2 3.2 2.7 RP6 DC 0.5 1 0.35 411 2.4 811 3.6 2.0 4.9 2.4 RP7 DC 3 6 2.1 345 4.6 87 4.2 3.8 4.1 1.1 RP8 DC 3 6 2.1 345 3.0 3 1.8 1.5 3.2 2.1 RP9 DC 3 6 2.1 345 5.2 148 5.5 5.1 7.8 1.5 RP10 DC 3 6 2.1 345 4.8 222 - 3.7 4.2 1.1 RP11 DC 3 6 2.1 360 4.2 15 - 2.4 5.4 2.3 RP12 DC 3 6 2.1 372 6.8 702 8.2 8.3 8.8 1.1 RP13 DC 3 6 2.1 376 5.7 1027 5.7 7.3 5.9 0.8 RP14 DC 19.5 36 14.6 350 10 214 5.8 5.7 5.3 0.9 RP15 LC creep 3 6 2.1 364 6.0 68 - 6.5 19.1 2.9 RP16 LC creep 3 6 2.1 385 3.8 43 - 2.5 14.0 5.6 RP17 LC cyclic 3 6 2.1 407 4.6 49 - 2.5 16.4 6.5 RP18 LC cyclic 3 6 2.1 408 4.0 40 - 2.5 14.2 5.7 RP19 LC cyclic 3 6 2.1 412 6.4 53 - 3.1 1.6 0.5 Fig. 5.3 shows the load-CMOD records for the 0.5m x 1m DC experiments (Fig. 5.3a), 3m x 6m DC experiments (Fig. 5.3b), 19.5m x 36 m DC experiment (Fig. 5.3c), and the 3m x 6m LC experiments (Fig. 5.3d). The area under the load-CMOD curve changed with size, rate, and loading type. The area under the monotonic ramp up to maximum load and up to complete fracture was used to compute J and W , Q f 106 Fracture Energy of Columnar Freshwater Ice: Influence of Loading Type, Loading Rate and Size [PIII] (a) (b) 0 500 1000 1500 2000 (c) (d) Figure 5.3. Load-CMOD records for the (a) 0.5m x 1m DC (b) 3m x 6m DC, (c) 19.5m x 36m DC and (d) 3m x 6m LC experiments [PIII]. 0 1 2 3 4 10 10 10 10 10 Figure 5.4. Variation of J and G as a function of the time to failure for the DC experiments [PIII]. Q VFCM 107 Fracture Energy of Columnar Freshwater Ice: Influence of Loading Type, Loading Rate and Size [PIII] 16 10 0 3 0 1 2 3 4 0 1 2 3 4 10 10 10 10 10 10 10 10 10 10 (a) (b) Figure 5.5. (a) The apparent fracture energy at crack growth initiation (J ) and (b) the work-of-fracture fracture energy (W ) as a function of the time to failure for the DC experiments. In (a), first-order power-law fits were applied separately to the data for the larger specimen (3m x 6m and 19.5m x 36m) and for the smaller specimen (0.5m x 1m) [PIII]. respectively (see Fig. 5.1). It is important to note that the area used here for evaluating J differs from the area defined in Fig. A1.2 of ASTM E1820-20b AST (2020) and used for the J calculation of metallic materials (Fig. 5.2). The used area (Figs. 5.1c and 5.1d) matches the area applied for other quasi brittle materials such as concrete and ceramics (Mindess et al., 1977; Homeny et al., 1980). There is a remarkable difference between the load-CMOD plots measured for columnar freshwater ice and quasi-brittle materials in general and those that typically occur for metallic materials. For metals, yielding occurs, followed by strain hardening then complete fracture (Fig. 5.2). On the other hand, yielding and strain hardening do not happen for quasi-brittle materials (Fig. 5.1). Fig. 5.4 shows the apparent fracture energies at crack growth initiation J and G as a function of the time to failure for the DC experiments. Interestingly, VFCM good agreement was obtained between J and G , especially for the 3m x 6m and Q VFCM 19.5m x 36m specimens. Although J and G are computed with different inherent Q VFCM assumptions, they generated similar results. J assumes a nonlinear elastic material behavior to characterize the crack tip deformation. G is based on cohesive crack VFCM mechanics coupled with a linear viscoelastic bulk behavior. This indicates that the fracture energy of a deeply cracked ECRP can be analyzed by two methods: Rice J-integral approximation and VFCM back-calculation. Fig. 5.5 shows the apparent fracture energy at crack growth initiation (J ) and the work-of-fracture (W ) as a function of the time to failure for the DC experiments. The J data indicates interrelated size and rate effects. There is a size effect at long failure times, but there is no size effect for fast experiments. The results from the larger plates (3m x 6m and 19.5m x 36m) are interchangeable and higher than the results from the smallest specimen (0.5m x 1m). For the larger specimens, J is 108 Fracture Energy of Columnar Freshwater Ice: Influence of Loading Type, Loading Rate and Size [PIII] decreasing with faster time to failure, but no significant rate effect was observed for the smallest specimen. The J − t relations are non-linear and can be well described Q f with a power-law relation. Fig. 5.5b displays similar size and rate effects as J ,but less clearly due to more scatter displayed by the values. Fig. 5.6 shows the J and W for the 3m x 6m DC and LC experiments as a function Q f of the time to failure. No clear effect of the creep-recovery loading on J was observed. However, the creep-recovery loading caused a significant increase in W for all the LC experiments excluding RP19. This can be seen from Figs. 5.3b and 5.3d. The LC experiments displayed a time-dependent behavior characterized by the gradual post-peak decline and large area under the load-CMOD curve. Therefore, while the creep and cyclic sequences did not affect the fracture energy at crack growth initiation (J ), they had a more pronounced effect on the work-of-fracture (W ). Q f 8 10 0 0 0 1 2 3 4 0 1 2 3 4 10 10 10 10 10 10 10 10 10 10 (a) (b) Figure 5.6. (a) The apparent fracture energy at crack growth initiation (J ) and (b) the work-of-fracture fracture energy (W ) as a function of the time to failure for the 3m x 6m DC and LC experiments. First-order power-law fits were applied to the data for the DC experiments. Note in (a), two LC experiments RP16 and RP17 gave the same J value [PIII]. Fig. 5.7 displays the fracture energy ratio (W /J ) as a function of the time to f Q failure. The W /J values for the DC experiments lay in a narrow range of 1 − 3. f Q The larger Experiment sizes showed a rate effect: At low rates, the ratio was 1, but it increased at high rates. The 0.5m x 1m experiments gave a higher ratio with no rate effect. Fig. 5.7 reveals also the effect of loading type. LC experiments RP16, RP17, and RP18 resulted in high W /J values. However, RP15 and RP19 were almost in the f Q same range as the DC experiments. This shows that the creep-recovery loading can increase the W /J in some cases. This leads to the conclusion that in the absence of f Q any creep or cyclic loading, the W of columnar freshwater S2 ice ranges between J f Q and 3J . 109 Fracture Energy of Columnar Freshwater Ice: Influence of Loading Type, Loading Rate and Size [PIII] -1 0 1 2 3 4 10 10 10 10 10 10 Figure 5.7. Fracture energy ratio W /J as a function of the time to failure for the DC and LC experi- f Q ments [PIII]. 110 6. Conclusion The thesis examined the fracture behavior of warm and floating columnar freshwater S2 ice under different loading (monotonic, creep, cyclic) scenarios, by varying the specimen size and loading rate/type in an in-situ experimental program. Chapter 2 introduced the reader to the ice growth procedure, the ice characterization, and the experimental setup. Chapter 3 covered the first part of the experimental program and examined the fracture behavior of warm columnar freshwater S2 ice under monotonic loading. Chapter 4 covered the second part of the experimental program and examined the fracture behavior of columnar freshwater ice under creep/cyclic- recovery and monotonic loading. Chapter 5 discussed the effect of loading rate and type and scale on the fracture energy of all the conducted experiments. The next sections summarize the main conclusions of each chapter and state potential future work. The text has been adapted from the author’s corresponding publications. 6.1 Grown Ice and Experimental Setup [PI, PII and PIII] Large-scale in-situ laboratory mode I fracture experiments of floating columnar (S2) freshwater ice were completed at the Ice Tank of Aalto University. The ice was very warm; the top surface was slightly cooler than -0.3 C. The ice thickness was very homogeneous throughout the ice sheet and increased from 34 cm to 41 cm during the test program that lasted four weeks. The specimen used was the edge-cracked rectangular plates (ECRP) loaded at the crack mouth up to fracture. The fracture mode was transgranular in all the experiments. 111 Conclusion 6.2 Fracture of Warm S2 Columnar Freshwater Ice Under Monotonic Loading: Size and Rate Effects [PI] 6.2.1 Summary Fourteen experiments on size (scale) and rate effects on the fracture behavior of warm, columnar (S2), freshwater ice were completed. The ECRP covered a size range of 1 : 39; the largest sample had dimensions of 19.5m x 36m. This size and the size range are to my knowledge the largest to any material under laboratory conditions. The loading rates applied led to loading durations from fewer than 2 seconds to more than 1000 seconds leading to an elastic response at the highest rates and a viscoelastic response at the lower rates. The experiments were displacement controlled (DC) and monotonically loaded up to fracture. The experiments were analysed by using the linear-elastic fracture mechanics (LEFM) and the non-linear viscoelastic fictitious crack model (VFCM) approach. For the LEFM analysis, expressions for the apparent fracture toughness (K ) and crack opening displacement were derived by following the weight function method by Dempsey and Mu (2014). For the VFCM, the approach formulated by Mulmule and Dempsey (1997) was adopted and combined with the Nelder-Mead optimization scheme (Nelder and Mead, 1965) to back-calculate the constitutive parameters: stress- separation curve, fracture energy, creep compliance constant, and size of the fracture process zone (PZ). The main observation from the experiments is that the size and rate effects were interrelated. There was a size effect at low loading rates, but no size effect was observed at the higher rates. This applied to the crack opening displacement near the crack tip (NCOD1) as measured, to K as an LEFM parameter, as well as to the stress-separation curve, the fracture energy (G ) and PZ as VFCM parameters. The ac rate dependent size effect is a novel result for any type of ice. Earlier both rate effects and size effects have been measured, but not how these two are related. Similar to earlier studies, the rate effects – when observed – followed power law type relations. The rate dependent size effect can be related with the fracture process at the vicinity of the crack tip: Both the measured NCOD1 and the calculated PZ of the larger specimen decrease with increasing rate and approach the values for the smaller specimen. As was illustrated by using the VFCM, the ice studied can be considered elastic only at the highest rates applied, the rates where the size effect vanishes. The measured size and rate effects can be expressed also as size dependent rate effects. For the larger specimen (3m x 6m and 19.5m x 36m), a clear rate effect was observed, while for the small specimen (0.5m x 1m) the rate effect was weak or absent. 112 Conclusion The results of the two larger, 3m x 6m and 19.5m x 36m, samples were interchange- able suggesting that the 3m x 6m sample size is large enough to give size-independent fracture results for this type of warm ice. From the experimental results and analysis presented, the following requirements are recommended. L/d ≥ 460 enough for polycrystalline homogeneity, av σ /σ < 0.2 notch sensitivity, n t in which L is the specimen size in the cracking direction and d is a measure of the av average grain size. The crack length must be selected to optimize the brittleness of the test plate. This study proved that the viscoelastic fictitious crack model is successful in treating the fracture of S2 columnar freshwater ice. 6.2.2 Research Suggestions • In future studies, it would be useful to do similar experiments but scale the ice- loading device contact length (D) with the specimen size (L, Fig. 3.1). It would be interesting to see if this scaling will affect the measured and the observed results. • Another interesting topic is the effect of the length of the crack front (i.e. plate thickness) on the fracture behavior. For metals, it is known that the state of stress near the tip of a crack tends to change from plane strain to plane stress as the ratio of width-to-thickness of a plate increases (Broek, 1974). However, it was proven for other quasi-brittle materials such as concrete that the thickness is not an important variable (Hillerborg, 1983; Mindess and Nadeau, 1976). This implies that there is no validity to the concepts of triaxiality and plane stress and plane strain idealizations in the fracture of quasi-brittle materials. It is believed that ice exhibits similar behavior as other quasi-brittle materials because a triaxial state of stress does not materialize. The tensile stress parallel to the crack front is relieved by microcracking parallel to the top and bottom surfaces or by creep, the latter taking place rapidly. However, it is important to prove this idea conclusively in an experimental program which vary the thickness of the specimen. To the author’s knowledge, these kind of experiments have not been conducted for ice. • The measured, LEFM and VFCM results showed a size dependent rate effect. There was a rate effect for the large specimens (3m x 6m and 19.5m x 36m) but weak or no rate effect for the smallest specimen (0.5m x 1m). Fig. 3.7b supported this conclusion further and portrayed that the rate effect is decreasing with decreasing specimen size and appears to disappear if specimen smaller than 0.5m x 1m were 113 Conclusion studied. It would be good to conduct experiments with specimens smaller than the smallest size tested here (0.5m x 1m) in order to confirm the lack of rate effect for small specimens. • The experiments suggested that a crack-parallel specimen size of about 460 d is av enough to achieve polycrystalline homogeneity, while 77 d is not. Finding the av exact requirement necessitates the testing of sample sizes which are larger than 0.5m x 1m and smaller than 3m x 6m. 6.3 Fracture of S2 Columnar Freshwater Ice Under Creep/Cyclic-Recovery Loading [PII] 6.3.1 Summary Five 3m x 6m warm freshwater S2 ice specimens were tested under creep/cyclic- recovery sequences followed by a monotonic ramp. The experiments were load controlled (LC) and led to complete fracture of the specimen. The purpose of this study was to examine the time-dependent behavior of freshwater ice using a joint experimental-modeling approach. In the experimental part, the experiments aimed to (1) measure and examine the time- dependent response of columnar freshwater S2 ice through the applied creep/cyclic- recovery sequences and (2) investigate the effect of creep and cyclic sequences on the fracture parameters/behavior through the fracture monotonic ramp. The current experiments were compared with the monotonically loaded DC experiments. The results showed that the creep and cyclic sequences had no clear effect on the failure load and the crack opening displacements at crack growth initiation. The ice response at the testing conditions was overall elastic-viscoplastic. The loading phases displayed an instantaneous transformation from the primary (transient) stage to the steady-state regime, which resulted in permanent (unrecoverable) displacement. The conducted experiments provided a novel observation for the time-dependent behavior of fresh- water ice. Though the delayed elastic component has been reported as a major creep component in freshwater ice, no significant viscoelasticity was detected in this study. Several factors were discussed as possibly contributing to the observed behavior: the very warm columnar freshwater ice, liquidity on the grain boundary, large sample size, coarse grain size, and nonlinear mechanisms in the fracture zone. Testing the effect of each factor on the ice response requires a different set of experiments that varies this 114 Conclusion factor only while keeping the other conditions fixed. In the modeling part, Schapery’s nonlinear constitutive model was applied for the displacement response at the crack mouth. The elastic-viscoplastic formulation generated a good fit with the experimental response of columnar freshwater S2 ice over the applied loading profile up to crack growth initiation, especially for the cyclic- recovery experiments. The model parameters were obtained via an optimization procedure using the N-M method by comparing the model and experimental CMOD values. The proposed model parameters are valid only for the studied ice type, geometry, specimen size, ice temperature, and the range of applied load experienced in the experiments. Schapery’s model was selected in this study, as it is able to capture the sort of time dependent behavior known to occur in ice and produces a simple and expedient way to help understand the observed behavior. 6.3.2 Research Suggestions • More thorough analysis of the experiments can be done with a physically-based approach that can give insight about the dislocation density, grain boundary sliding and the operating mechanisms. • It would be useful to try longer recovery period to see if this will affect the accumu- lation of unrecoverable deformation. • The results showed no effect of the creep/cyclic-recovery loading on the fracture properties: the failure load and the crack opening displacements at crack growth initiation. It would be interesting to try higher creep and cyclic load levels for similar testing conditions and observe again the effects on the fracture properties. • The current experiments suggest that for the large sample size and the kind of ice studied (very warm freshwater ice) under the loading applied, viscoelasticity is not an important deformation component. The experimental results support this prediction; but more tests, that vary the temperature and size, are needed to make more general conclusions. 115 Conclusion 6.4 Fracture Energy of Columnar Freshwater Ice: Influence of Loading Type, Loading Rate and Size [PIII] 6.4.1 Summary The mode I fracture energies of columnar freshwater S2 ice were quantified using different methods and the influence of the type of loading, test size and loading rate on the fracture energy was investigated. The apparent fracture energy at crack growth initiation was computed in two ways: 1) J , via Rice’s J-integral expression for deeply-cracked edge-cracked plates using the load-CMOD record and 2) G , via VFCM a viscoelastic fictitious crack model using the back-calculated stress-separation curve. The work-of-fracture (W ) was obtained from the load-CMOD record. Values of J , f Q G , and W were analyzed considering the effects of scale and loading rate and VFCM f type. A good match was obtained between the J and G values, especially for the Q VFCM 3m x 6m and 19.5m x 36m specimens. J displayed a size dependent rate effect and a rate dependent size effect. There was a size effect at low loading rates, but no size effect was observed at the higher rates. For the larger specimens (3m x 6m and 19.5m x 36m), a clear rate effect was observed, while for the small specimen (0.5m x 1m) the rate effect was almost absent. W showed similar rate and size effects. The results of the DC experiments were compared against the LC experiments to indicate the effect of creep-recovery loading on the fracture energy. No clear effect of the creep and cyclic sequences on J was detected. However, the creep-recovery loading caused a noticeable increase in W for most of the LC experiments. The W /J values for the DC experiments lay in a narrow range (1-3). The larger f Q experiments (3m x 6m and 19.5m x 36m) gave a ratio of 1 at low rates and a higher ratio at higher rates. For the 0.5m x 1m experiments, the ratio was rate independent and was higher than 1. The creep-recovery loading led to higher W /J in some cases. f Q This study proves the significance of the deep-notched edge-cracked rectangular plate geometry. Several ways of analysis can be applied; Rice J-integral and VFCM were implemented with similar results. 6.4.2 Research Suggestions The experiments showed an interesting similarity between the J and G values. Q VFCM It is important to test this observation by conducting more experiments using different sample sizes and testing conditions. 116 A. Weight Function For An Edge-Cracked Rectangular Plate The weight function method is used in fracture mechanics to determine the expressions of the stress intensity factor and the crack opening displacement [Wu and Carlsson, 1991]. Knowledge of a two-dimensional elastic crack solution (the reference solution) as a function of crack length A for any reference crack face pressure Σ (X) enables the 0 r determination of the stress intensity factor K(A ) and crack opening displacement U(A , X) for the same body under arbitrary loading Σ(X): The derivation of the weight function for an ECRP was given by Dempsey and Mu (2014). E ∂U K(A ) = Σ(X)H (A , X)dX, H (A , X) = (A , X) 0 r 0 r 0 0 K (A ) ∂A 0 r 0 0 E U(A , X) = K(T)H (T, X)dT (A1) 0 r in which H (A , X) denotes the weight function (valid for X ≤ A ) and E is r 0 0 the elastic modulus. The coordinate X is along the crack line with its origin at the crack mouth (see Fig. 2.4a, Chapter 2). K (A ) is the known or reference stress r 0 intensity factor while 2U (A , X) represents the known or reference crack face opening r 0 dis- placement.Chapter 2 introduced the reader to the ice growth procedure, the ice characterization, and the experimental setup. Chapter 3 covered the first part of the experimental program and examined the fracture behavior of warm columnar freshwater S2 ice under monotonic loading. Chapter 4 covered the second part of the experimental program The final expression for the weight function is 1 x 3 i− h (a , x) = √ G (a )(1 − ) (A2) r 0 i 0 2πa i=1 117 Weight Function For An Edge-Cracked Rectangular Plate The G (i = 1, 2, 3, 4, 5) functions are given by G (a ) = 2.0 1 0 6 105E(a ) aF (a ) 0 0 G (a ) = + √ + 4 − 30 2 0 (1 − a ) F(a ) 0 1 − a 0 S (a ) 26 455E(a ) 52a F (a ) 3 0 0 0 0 G (a ) = − − √ − + 86 3 0 3/2 (1 − a ) (1 − a ) 3F(a ) 1 − a 0 0 0 S (a ) 154 539E(a ) 308a F (a ) 434 4 0 0 0 0 G (a ) = + + √ + − 4 0 3/2 5(1 − a ) 15F(a ) 5 (1 − a ) 1 − a 0 0 0 S (a ) 54 189E(a ) 36a F (a ) 144 5 0 0 0 0 G (a ) = − − − + (A3) 5 0 √ 3/2 5(1 − a ) 5F(a ) 5 (1 − a ) 0 0 1 − a 0 where E(a ) = [3πΦ(a ) − V(a )] /8 2F(a ) 0 0 0 0 35 V(a ) S (a ) = 3πF(a ) − 3 0 0 F(a ) 4 2 0 7 V(a ) S (a ) = √ −12πF(a ) + 5 4 0 0 F(a ) 2 2 0 9 V(a ) S (a ) = √ 7πF(a ) − 3 5 0 0 F(a ) 4 2 V(a ) = (1 + a )V(a ) + a (1 − a )V (a ) (A4) 0 0 0 0 0 0 The ( ) denotes differentiation with repect to a . Functions F(a ), V(a ), and Φ(a ) 0 0 0 0 are related to the reference SIF K (a ), the reference CMOD V (a ), and the reference r 0 r 0 normalized crack opening area φ(a ), respectively, by F(a ) V(a ) Φ(a ) 0 0 0 K (A ) = σ πA , V (a ) = ,φ(a ) = (A5) r 0 0 r 0 0 2 2 (1 − a ) (1 − a ) (1 − a ) 0 0 where σ is the reference uniform crack face pressure. The actual expressions for F(a ), V(a ), and Φ(a ) were fitted with a seventh-order 0 0 0 polynomial as follows 7 7 7 i i i F(a ) = α s , V(a ) = γ s , Φ(a ) = ϕ s (A6) 0 i 0 i 0 i i=0 i=0 i=0 where the coefficients α , γ „ and ϕ (i = 0, 1,... 7), consecutively, were found to be i i i α:1.12150, −1.59550, 6.83570, −16.4318, 26.2079, −26.3156, 14.8991, −3.59980 γ:2.90860, −5.56710, 19.3629, −38.1097, 56.5619, −52.6474, 27.4002, −5.9579 ϕ:0.62890, −1.20010, 4.1109, −8.6992, 12.7842, −11.6705, 5.8923, −1.2176 (A7) 118 Bibliography 1. Santaoja K. Ph.D. Thesis: Mathematical modelling of deformation mechanisms in ice. Technical Research Centre of Finland, Espoo; 1990. 2. ASTM Standard, E1820-20b, Standard test method for measurement of fracture toughness. ASTM International, 100 Barr Harbor Drive, PO Box C700, West Conshohocken, PA 19428-2959. USA; 2020. 3. Kirsch C. Die theorie der elastizitat und die bedurfnisse der festigkeitslehre. Zeitschrift des Vereines Deutscher Ingenieure 1898;42:797–807. 4. Inglis C. Stresses in a plate due to the presence of cracks and sharp corners. Transactions of the Institute of Naval Architects 1913;55:219–41. 5. Griffith AA. The phenomena of rupture and flow in solids. Philosophical transactions of the Royal Society of London Series A 1921;221:163–98. 6. Irwin G. Fracture dynamics. In: Fracturing of Metals. 1948:147–66. 7. Orowan E. Fracture and strength of solids. Reports on Progress in Physics 1949;12:185–232. 8. Westergaard HM. Bearing pressures and cracks. Journal of Applied Mechanics 1939;6:49–53. 9. Irwin GR. Analysis of stresses and strains near the end of a crack transversing a plate. Journal of Applied Mechanics 1957;24:361–4. 10. Rice JR. Plastic yielding at a crack tip. In: Proceedings of ICF1, Japan; vol. 1. 1965:283–308. 11. Elices M, Guinea GV, Gomez J, Planas J. The cohesive zone model: advantages, limitations and challenges. Engineering Fracture Mechanics 2002;69:137–63. 12. Dempsey J, Cole D, Wang S. Tensile fracture of a single crack in first-year sea ice. Philosoph- ical Transactions of the Royal Society A: Mathematical, Physical and Engineering Sciences 2018;376(2129):20170346. 13. Bazant ZP, Planas J. Fracture and Size Effect in Concrete and Other Quasibrittle Materials. CRC Press; 1998. 14. Bilby BA, Cottrell AH, Swinden K. The spread of plastic yield from a notch. Proceedings of the Royal Society of London 1963;A272:304–14. 15. Smith E. The structure in the vicinity of a crack tip: a general theory based on the cohesive zone model. Engineering Fracture Mechanics 1974;6:213–22. 16. Dugdale DS. Yielding of steel sheets containing slits. Journal of the Mechanics and Physics of Solids 1960;8:100–4. 119 Bibliography 17. Barenblatt GI. The mathematical theory of equilibrium cracks in brittle fracture. Advances in Applied Mechanics 1962;7:55–129. 18. Hillerborg A, Modéer M, Petersson PE. Analysis of crack formation and crack growth in concrete by means of fracture mechanics and finite elements. Cement and concrete research 1976;6:773–81. 19. Van Mier J, Van Vliet M. Uniaxial tension test for the determination of fracture parameters of concrete: state of the art. Engineering Fracture Mechanics 2002;69:235–47. 20. Mulmule S, Dempsey JP. A viscoelastic fictitious crack model for the fracture of sea ice. Mechanics of Time-Dependent Materials 1997;1:331–56. 21. Michel B, Ramseier RO. Classification of river and lake ice. Canadian Geotechnical Journal 1971;8:36–45. 22. Shestov A. Birefringence in ice crystals. Principles and application in sea ice microstructure studiese. 24th IAHR Int Symp on Ice, 2018;Russia. 23. Liu H, Loop W. Fracture toughness of freshwater ice. CRREL Technical Note. USA Cold Regions Research and Engineering Laboratory; 1972. 24. Goodman D, Tabor D. Fracture toughness of ice: a preliminary account of some new experiments. Journal of Glaciology 1978;21:651–60. 25. Liu H, Miller K. Fracture toughness of fresh-water ice. Journal of Glaciology 1979;22:135–43. 26. Hamza H, Muggeridge D. Plane strain fracture toughness (K ) of fresh water ice. In: 1c International Conference on Port and Ocean Engineering under Arctic Conditions; vol. 1. 1979:697–707. 27. Urabe N, Yoshitake A. Strain rate dependent fracture toughness (K ) of pure ice and sea 1C ice. In: Proceedings of the 6th IAHR Symposium on Ice, Quebec City, PQ, Canada; vol. 2. 1981:551–63. 28. Timco G, Frederking R. Comparative strengths of fresh water ice. Cold Regions Science and Technology 1982;6:21–7. 29. Danilenko V. Determination of crack resistance (K ) of freshwater ice. Mechanics of Solids 1c 1985;20:131–6. 30. Parsons B, Snellen J. Fracture toughness of fresh water prototype ice and carbamide model ice. In: 8th International Conference on Port and Ocean Engineering under Arctic Conditions, Narsarsuaq, Greenland; vol. 1. 1985:128–37. 31. Timco GW, Frederking RMW. The effects of anisotropy and microcracks on the fracture tough- ness (K ) of freshwater ice. In: Proceedings of the 5th International OMAE Symposium, 1c Tokyo, Japan; vol. 4. 1986:341–8. 32. Nixon W, Schulson E. Fracture toughness of freshwater ice as a function of loading rate. In: Proceedings of the 1st International Conference on Ice Technology, Cambridge, MA. 1986:287–96. 33. Tuhkuri J. The applicability of LEFM and the fracture toughness (K ) to sea ice. In: The IC 9th International Conference on Port and Ocean Engineering under Arctic Conditions, Fairbanks, Alaska; vol. 1. 1987:21–32. 120 Bibliography 34. Bentley DL, Dempsey JP, Sodhi DS, Wei Y. Fracture of S2 columnar freshwater ice: floating double cantilever beam tests. In: Proceedings of the 9th International IAHR Symposium on Ice, Sapporo, Japan; vol. 1. 1988:152–61. 35. Dempsey JP, Wei Y, DeFranco S, Ruben R, Frachetti R. Fracture toughness of S2 columnar freshwater ice: crack length and specimen size effects – part II. In: The 10th International Conference on Port and Ocean Engineering under Arctic Conditions, Luleå, Sweden; vol. 1. 1989:199–207. 36. Parsons B, Snellen J, Muggeridge D. The double torsion test applied to fine grained freshwater columnar ice and sea ice. Mechanics of Creep Brittle Materials 1989;1:188–200. 37. Wei Y, DeFranco S, Dempsey JP. Crack-fabrication techniques and their effects on the fracture toughness and CTOD for fresh-water columnar ice. Journal of Glaciology 1991;37:270–80. 38. Weber L, Nixon W. Fracture toughness of freshwater ice – Part I: Experimental technique and results. Journal of Offshore Mechanics and Arctic Engineering 1996a;118:135–40. 39. Weber L, Nixon W. Fracture toughness of freshwater ice – Part II: Analysis and micrography. Journal of Offshore Mechanics and Arctic Engineering 1996b;118:141–7. 40. Xu X, Jeronimidis G, Atkins A, Trusty P. Rate-dependent fracture toughness of pure polycrys- talline ice. Journal of Materials Science 2004;39:225–33. 41. Weeks WF, Assur A. Fracture of lake and sea ice. CRREL Research Report 269. USA Cold Regions Research and Engineering Laboratory; 1969. 42. Dempsey JP. The fracture toughness of ice. In: Ice-Structure Interaction. Springer, Berlin, Heidelberg; 1991:109–45. 43. Dempsey J, Adamson R, Mulmule S. Scale effects on the in-situ tensile strength and fracture of ice. Part II: First-year sea ice at Resolute, N.W.T. International Journal of Fracture 1999a;95:347–66. 44. Dempsey JP, Wei Y, DeFranco SJ. Notch sensitivity and brittleness in fracture testing of s2 columnar freshwater ice. International Journal of Fracture 1992;53:101–20. 45. Mulmule SV, Dempsey JP. Scale effects on sea ice fracture. Mechanics of Cohesive-Frictional Materials 1999;4:505–24. 46. Carpinteri A. Notch sensitivity in fracture testing of aggregative materials. Engineering Fracture Mechanics 1982;16:467–81. 47. Abdel-Tawab K, Rodin GJ. On the relevance of linear elastic fracture mechanics to ice. International Journal of Fracture 1993;62(2):171–81. 48. Abdel-Tawab K, Rodin GJ. Fracture size effects and polycrystalline inhomogeneity. Interna- tional journal of fracture 1998;93(1-4):247–59. 49. Dempsey JP, DeFranco SJ, Adamson RM, Mulmule SV. Scale effects on the in-situ tensile strength and fracture of ice. Part I: Large grained freshwater ice at Spray Lakes Reservoir, Alberta. International Journal of Fracture 1999b;95:325–45. 50. Mulmule S, Dempsey J. LEFM size requirements for the fracture testing of sea ice. Interna- tional journal of fracture 2000;102(1):85–98. 51. Schwarz J, Frederking R, Gavrillo V, Petrov I, Hirayama KI, Mellor M, Tryde P, Vaudrey K. Standardized testing methods for measuring mechanical properties of ice. Cold Regions Science and Technology 1981;4(3):245–53. 121 Bibliography 52. Jellinek H, Brill R. Viscoelastic properties of ice. Journal of Applied Physics 1956;27(10):1198–209. 53. Sinha NK. Rheology of columnar-grained ice. Experimental Mechanics 1978;18:464–70. 54. Michel B. A mechanical model of creep of polycrystalline ice. Canadian Geotechnical Journal 1978a;15:155–70. 55. Le Gac H, Duval P. Constitutive relations for the non elastic deformation of polycrystalline ice. In: Physics and Mechanics of Ice. 1980:51–9. 56. Ashby MF, Duval P. The creep of polycrystalline ice. Cold Regions Science and Technology 1985;11:285–300. 57. Sunder SS, Wu MS. A differential flow model for polycrystalline ice. Cold Regions Science and Technology 1989;16:45–62. 58. Mellor M, Cole DM. Stress/strain/time relations for ice under uniaxial compression. Cold Regions Science and Technology 1983;6:207–30. 59. Cole DM. Reversed direct-stress testing of ice: Initial experimental results and analysis. Cold Regions Science and Technology 1990;18:303–21. 60. Duval P, Kalifa P, Meyssonnier J. Creep constitutive equations for polycrystalline ice and effect of microcracking. In: International Union on Theoretical and Applied Mechanics (IUTAM). 1991:55–67. 61. Sunder SS, Wu MS. On the constitutive modeling of transient creep in polycrystalline ice. Cold Regions Science and Technology 1990;18:267–94. 62. Abdel-Tawab K, Rodin GJ. Analysis of primary creep of S2 fresh-water and saline ice. Cold Regions Science and Technology 1997;26:83–96. 63. Schapery RA. On the characterization of nonlinear viscoelastic materials. Polymer Engineering & Science 1969;9:295–310. 64. Murdza A, Schulson EM, Renshaw CE. Strengthening of columnar-grained freshwater ice through cyclic flexural loading. Journal of Glaciology 2020;:1–11. 65. Iliescu D, Murdza A, Schulson EM, Renshaw CE. Strengthening ice through cyclic loading. Journal of Glaciology 2017;63(240):663–9. 66. Iliescu D, Schulson E. Brittle compressive failure of ice: monotonic versus cyclic loading. Acta materialia 2002;50(8):2163–72. 67. Jorgen VG, Picu BC. Effect of step-loading history and related grain-boundary fatigue in freshwater columnar ice in the brittle deformation regime. Philosophical magazine letters 1998;77(5):241–7. 68. Ashton GD. River and Lake Ice Engineering. Water Resources Publication, Littletown, Colorado; 1986. 69. Gow AJ, Ueda HT, Govoni JW, Kalafut J. Temperature and structure dependence of the flexural strength and modulus of freshwater model ice; vol. 88-6. USA Cold Regions Research and Engineering Laboratory; 1988. 70. Cole DM. Effect of grain size on the internal fracturing of polycrystalline ice. CRREL Report 86-5. USA Cold Regions Research and Engineering Laboratory; 1986. 122 Bibliography 71. Kamb WB. Refraction corrections for universal stage measurements. I. uniaxial crystals. American Mineralogist: Journal of Earth and Planetary Materials 1962;47:227–45. 72. Langway CC. Ice fabrics and the universal stage. CRREL Technical Report 62. USA Cold Regions Research and Engineering Laboratory; 1958. 73. Dempsey JP, Mu Z. Weight function for an edge-cracked rectangular plate. Engineering Fracture Mechanics 2014;132:93–103. 74. Lu W, Løset S, Shestov A, Lubbad R. Design of a field test for measuring the fracture toughness of sea ice. In: Proceedings of the 23rd International Conference on Port and Ocean Engineering Under Arctic Conditions, Trondheim, Norway. 2015:. 75. Bossuyt S. Optimized patterns for digital image correlation. In: Imaging Methods for Novel Materials and Challenging Applications; vol. 3. Springer; 2013:239–48. 76. Ahmad W. Digital Image Correlation Analysis of Ice Fracture Experiments. Aalto University (master thesis); 2019. 77. Cole DM. The effect of creep on the constitutive behavior of saline ice at low strains. In: Ice Mechanics-1993, Proceedings 1st Joint Mechanics Meeting of ASME. ASCE and SES; vol. 163. ASMEAMD; 1993:261–72. 78. Schapery RA. Viscoelastic deformation behavior of ice based on micromechanical models. In: Ice Mechanics-1993; vol. 163. ASME AMD; 1993:15–33. 79. Nelder JA, Mead R. A simplex method for function minimization. The Computer Journal 1965;7:308–13. 80. Carter D. Ph.D. Thesis: Lois et mecanismes de l’apparente fracture fragile de la glace de riviere et de lac. Universite Laval, Section Mecanique des Glaces; 1971. 81. Gold LW. Engineering properties of fresh-water ice. Journal of Glaciology 1977;19:197–212. 82. Michel B. The strength of polycrystalline ice. Canadian Journal of Civil Engineering 1978b;5:285–300. 83. Bazant Z. Size effect in blunt fracture: Concrete, rock, metal. Journal of Engineering Mechanics 1984;110:518–35. 84. Dash J, Rempel A, Wettlaufer J. The physics of premelted ice and its geophysical consequences. Reviews of modern physics 2006;78(3):695. 85. Bell RL, Langdon TG. An investigation of grain-boundary sliding during creep. Journal of Materials Science 1967;2(4):313–23. 86. Hillerborg A. Analysis of one single crack. In: Fracture Mechanics of Concrete. Elsevier, Amsterdam; 1983:223–49. 87. Broek D. Elementary Engineering Fracture Mechanics. Noordhoff Publishing Co., Leiden; 88. Mindess S, Nadeau JS. Effect of notch width on (K ) for mortar and concrete. Cement and 1c Concrete Research 1976;6(4):529–34. 89. Riedel H, Rice J. Tensile cracks in creeping solids. In: Fracture Mechanics. Proceedings of the Twelfth National Symposium on Fracture Mechanics. ASTM STP 700; 1980:112–30. 123 Bibliography 90. Veerman C, Muller T. The location of the apparent rotation axis in notched bend testing. Engineering Fracture Mechanics 1972;4:25–32. 91. Rodin GJ. On fracture of warm ice. In: IUTAM Symposium on Scaling Laws in Ice Mechanics and Ice Dynamics. Springer; 2001:207–16. 92. Wang Y, Ballarini R, Rodin GJ. Crack-tip parameters in polycrystalline plates with soft grain boundaries. Journal of engineering mechanics 2008;134(1):100–9. 93. Schulson E, Duval P. Creep and Fracture of Ice. Cambridge University Press; 2009. 94. Schapery R. Thermoviscoelastic constitutive equations for polycrystalline ice. Journal of Cold Regions Engineering 1997;11:146–57. 95. Adamson RM, Dempsey JP. Field-scale in-situ compliance of arctic first-year sea ice. Journal of Cold Regions Engineering 1998;12:52–63. 96. LeClair ES, Schapery RA, Dempsey JP. A broad-spectrum constitutive modeling technique applied to saline ice. International Journal of Fracture 1999;97:209–26. 97. LeClair ES, Schapery RA, Dempsey JP. Tensile creep of saline ice. In: Symposium on Inelas- ticity and Damage in Solids Subject to Microstructural Change. St. John’s, Newfoundland, Canada; 1996:143–53. 98. Flügge W. Viscoelasticity. Springer-Verlag. Berlin; 1975. 99. Lou YC, Schapery RA. Viscoelastic characterization of a nonlinear fiber-reinforced plastic. Journal of Composite Materials 1971;5:208–34. 100. Papanicolaou G, Zaoutsos S, Cardon A. Further development of a data reduction method for the nonlinear viscoelastic characterization of frps. Composites Part A: Applied Science and Manufacturing 1999;30:839–48. 101. Mellor M, Cole D. Cyclic loading and fatigue in ice. Cold regions science and technology 1981;4(1):41–53. 102. Mellor M, Cole DM. Deformation and failure of ice under constant stress or constant strain-rate. Cold Regions Science and Technology 1982;5(3):201–19. 103. Duval P. Anelastic behaviour of polycrystalline ice. Journal of Glaciology 1978;21(85):621–8. 104. Sinha NK. Grain boundary sliding in polycrystalline materials. Philosophical Magazine A 1979;40(6):825–42. 105. Cole DM. A model for the anelastic straining of saline ice subjected to cyclic loading. Philosophical Magazine A 1995;72(1):231–48. 106. Cole D. On the physical basis for the creep of ice: the high temperature regime. Journal of Glaciology 2020;66(257):401–14. 107. Muto H, Sakai M. Grain-boundary sliding and grain interlocking in the creep deformation of two-phase ceramics. Journal of the American Ceramic Society 1998;81(6):1611–21. 108. Gasdaska CJ. Tensile creep in an in situ reinforced silicon nitride. Journal of the American Ceramic Society 1994;77(9):2408–18. 109. Knauss WG. A review of fracture in viscoelastic materials. International Journal of Fracture 2015;196(1-2):99–146. 124 Bibliography 110. Rice JR. A path independent integral and the approximate analysis of strain concentration by notches and cracks. Journal of Applied Mechanics 1968;35:379–86. 111. Mindess S, Lawrence FV, Kesler CE. The J-Integral as a fracture criterion for fiber reinforced concrete. Cement and Concrete Research 1977;7(6):731–42. 112. Homeny J, Darroudi T, Bradt RC. J-Integral measurements of the fracture of 50% alumina refractories. Journal of the American Ceramic Society 1980;63(5-6):326–31. 113. Rice JR, Paris PC, Merkle JG. Some further results of J-Integral analysis and estimates. In: Progress in flaw growth and fracture toughness testing.ASTM STP 536. American Society for Testing and Materials; 1973:231–45. 114. Begley J, Landes J. The J integral as a fracture criterion. In: Fracture Toughness, Proceedings of the 1971 National Symposium on Fracture Mechanics, Part II. ASTM STP 514, American Society for Testing and Materials; 1972:1–20. 115. Landes JD, Begley JA. The effect of specimen geometry on J . In: Fracture Toughness, Ic Proceedings of the 1971 National Symposium on Fracture Mechanics. ASTM STP 514, American Society for Testing and Materials; 1972:24–39. 116. Bucci RJ, Paris PC, Landes JD, Rice JR. J integral estimation procedures. In: Fracture Toughness, Proceedings of the 1971 National Symposium on Fracture Mechanics, Part 1I. ASTM STP 514, American Society for Testing and Materials; 1972:40–69. 117. Landes JD, Begley JA. Test results from J-integral studies: An attempt to establish a J Ic testing procedure. In: Fracture analysis: proceedings of the 1973 National symposium on fracture mechanics, Part II. ASTM STP 560, American Society for Testing and Materials; 1974:170–86. 118. Clarke GA, Andrews WR, Paris PC, Schmidt DW. Single specimen tests for J determina- Ic tion. In: Mechanics of crack growth. ASTM STP 590, American Society for Testing and Materials; 1976:27–42. 119. Logsdon W. Elastic plastic (J ) fracture toughness values: Their experimental determination Ic and comparison with conventional linear elastic (K ) fracture toughness values for five Ic materials. In: Mechanics of Crack Growth. ASTM STP 590, American Society for Testing and Materials; 1976:43–60. 120. Yoder GR, Griffis CA. Application of the J-integral to the initiation of crack extension in a titanium 6al-4v alloy. In: Mechanics of crack growth. ASTM STP 590, American Society for Testing and Materials; 1976:61–81. 121. Griffis CA, Yoder GR. Initial crack extension in two intermediate-strength aluminum alloys. Transactions ASME, Journal of Engineering Materials and Technology 1976;98:152–8. 122. Sakai M, Yoshimura JI, Goto Y, Inagaki M. R-curve behavior of a polycrystalline graphite: microcracking and grain bridging in the wake region. Journal of the American Ceramic society 1988;71(8):609–16. 123. Li F, Shen W, Sun X, Gong D. Method for determining elasto-plastic fracture toughness of ice (in Chinese). Journal of Glaciology and Geocryology 1996;18:178–83. 124. Nakayama J. A bending method for direct measurement of fracture energy of brittle material. Japanese Journal of Applied Physics 1964;3(7):422–3. 125. Tattersall HG, Tappin G. The work of fracture and its measurement in metals, ceramics and other materials. Journal of Materials Science 1966;1(3):296–301. 125 Bibliography 126. Bažant ZP. Analysis of work-of-fracture method for measuring fracture energy of concrete. Journal of Engineering Mechanics 1996;122(2):138–44. 127. Dempsey JP, Xie Y, Adamson RM, Farmer DM. Fracture of a ridged multi-year arctic sea ice floe. Cold Regions Science and Technology 2012;76:63–8. 128. Wu XR, Carlsson J. Weight functions and stress intensity factor solutions. Pergamon; 1991. 126 lto- 1 / 1 9 T G*e j + "Researcher a s A t alt o Univer t y hav e foun d stron g evidenc e tha w t arm ic e – tha t i,c e ver y clos e i n temperatur e t o zer o degree C s el u – s may fractur d eiff ent lty ha t n h k eind o i sc f te ypical ls ytudie id ln aboratories o r nature.. . Understandin g how ic e break scrucia flo e rnsurin g saf e harbour a snd bridg e ic s no o cllimate s a w ,se a ltls ransp tatio tn h ug h hist ically ice-hea v ry egion A s g.s lob w alarmi bri cs ha e t oo s nce- predictab ls eeason a clondition s th , r eule u snderp g frastructure engi erin ag r be ein tg est e a d cro s bs orde r as n cd onti n ts. 'W e nee d t o stud y warm ic e becaus e it 'w s ha w t e'r e seein g i n nature; glob a w l arm ih g sappe g T.h m e echa c a plropertie o s ic f a en h dow rie t spon d t fo s or c m e a b fy u e ndamental d ly iffere n w the i n tw 'sarm rath e th r col d a w t s ,r eaditio ls ly tu d i s t y,a ' y Im s El Gh mti .A . .w s m e crondition s ie ncreasing le yxpecte id n previous lf yr i d region ls ik t eh G e rea L take o s B r alt iS ce – a on o etfhe world s 'busies m t arin e area s – Pro fJ . ukk a Tuhkurs iay s ii s tcrucia tlo u ersta t h m eechan i w o cs far i m c o 'I e . u . f s ..h r i p as i rastructu rle ik b eridg an w d in t d urbin hav b eee d n igne fd or fair lp yredictab lse eason w s ,n eee td k ono w wh a htappen w she g nlobal warmi n bg rin g ns e c w diti lIo st.o k lis k t h e o e r ld ul e m s a y ht old up ,T ' uhkur i says.. . P ss leas e W : ar m ic m e a fy ractu diffe nt lty ha c no lid c e K ,atrina Jurva/Aalt U o versi ty BUSINESS + ECONOMY IS - - - -9 (print ) AR T + IS - - - -6 ( f) DESIGN + IS - (print ) ARCHITECTURE IS - ( f) SCIENCE + TECHNOLOGY A la t o Universi yt CROSSOVER Scho o o Elf gn ineeri gn Mecha in c aE l ng ni eer ni g DOCTORAL .aalto .fi DISSER TATIONS dd fa MF SH 202 43 DD aA www dp 2494 9971 NS de 4394 9971 NS dp 4350 46 259 879 NB de 3350 46 259 879 NB in er er er er on no no se se se fn dn dn dn gi ra ra ara na an na in nin ni en en ni nin ni gn gn gn ro or ro si re is si is
Materialia – Unpaywall
Published: Dec 1, 2021
You can share this free article with as many people as you like with the url below! We hope you enjoy this feature!
Read and print from thousands of top scholarly journals.
Already have an account? Log in
Bookmark this article. You can see your Bookmarks on your DeepDyve Library.
To save an article, log in first, or sign up for a DeepDyve account if you don’t already have one.
Copy and paste the desired citation format or use the link below to download a file formatted for EndNote
Access the full text.
Sign up today, get DeepDyve free for 14 days.
All DeepDyve websites use cookies to improve your online experience. They were placed on your computer when you launched this website. You can change your cookie settings through your browser.