1 - l . ‘ r . V . . v V V VV 2 VV . . . .. V: J, . .. . 3 . A ... . ..V n V . , xii V5. ..t... .V ! . . . V V V V V V I. . ..... . . V J ... . . ... . :VI4 . «f t.'..V!..,V .VIZ] . V . V . I V. V . V . . V . . . fro 1.5115113”! I «1. 1.. . , _ : , : . V . .V . , V V . . ... . .V V . ,. . .V V... r 4441.8}.{iisu1 15V , V ,3: — . . V V V . . . .. V . . ,. . _ f: ...; «lazf. rlr, , .4. ”fin. Humans Janna. V .V f, . VVVZ ,. . H {4.711 .1? :. MW.) 1 . . V 1hrv‘.j$la 4.11 o Int-15:» 4 , ”£425.94 V rig-Jr,” in! «initiatf .1p in“. .. V 9.1 .. , (fun ,Wrrf.» *. ...“. . V1 V.» V . . V. c . V , V . ., “.1 «4.5th.. Main? .V . V V V7)...‘a 5: ’43 x J I _ .fl, G v . V. . 59.. . V.. E. .. V . a! it.) V1.7 2.4%.an .V V. .1 r .l _ . )V. . v V hflf 1.55.2". A12 . V r V1H¢J4Tz ; finnéwfifg _ vrfifiwfifiulhfiunl t a. V V 1.. .2 Vuhfififir ...rxrsil 3va.!it1..r._.rh..!1. wuuhnfifln . V . . r (rite-Fm .Vc. . . . 44!!! V . “7;: . 47.159... :1..— )n V X: If! it.» . 14» V . V .llrtr. It; 4.. .. .HGHJWH . .JVrnnrflrfi r .fliJ-ly ..hN {1.89;}! x V V (.1 x I , ; WSJMWMfrith if} 1 V #3434 (lab Vtmvw J ....) ‘ 3.2.2 .‘hflwfiflflvir. figurfigflwwfifzgn V. ... VVV... .. V .25 Iliyzi‘lm in, . V V . V n . mm}!!! NH... xi!.fi§:vfiu - H .V. , BN4 aria: “10%.; v‘rvvljvuvfi :21 ( . ....D .. V V . 1%1fl§vafluLVV)le4t.:u , in! ....§A!1,I:S.9Lv.?5t# I I": .3 .211!;.11¥52#117. '01 I, V (L 1.3 , . iii‘A , a. V V A ~ 13.! V $311 w.‘. . I . I . V Lieu .. . V :3 a 1! ya. 1.1.1,“); 1.3 15“ .17le11. . V V n l V . 1‘. ...; A . V n I — :94 If . . .: V .. . il‘Yifil‘u -3. 22.4““? . v .Q 6.11 YtVJI-‘i.p .. .144. V 1:. (IV. 1 .. 3.513.145 . I‘VVu‘.:-V:Vx}hm V V V ..Ilvn}. WvTiwm 111V V, «ills»............V..IVH:1.:.1I?1V.. .V . V .V .V V . {dumervittannuuwluv ...ls «tllhfinrfll. I. 74.131254 3V 1.61.!!! {114.11. ~tEVl. . ~ . . . « .. .- T i V .91... .anJJ... ,...V V.VVVV.IVV:\u|V.K.ivanvut1v-J.. THESIS LIBRARY Michigan State University nr’ ABSTRACT NONLINEAR, QUASI—STATIC BEHAVIOR OF SOME PHOTOELASTIC AND MECHANICAL MODEL MATERIALS by Ernst W. Kiesling A study was made of the effects of time upon the birefringent and mechanical properties of four plastics: Polyester Resin CR-39 (Cast Optics Co., Inc.); Polyester Resin PS—l and Epoxy Resin PS—2 (Photolastic, Inc.); and Polyester Resin Palatal P6—K (Badische Anilin und Soda— Fabrik9 Germany)° The isothermal, quasi—static behavior under constant, uniaxial stress conditions was examined exhaustively—~for approximately five decades of logarithmic timeV3 and for a range of stress sufficiently broad to cover the linear and much of the nonlinear range (where non— linearities appeared)o The stresses limiting the linear range of behavior-— called linear limit stresses—wwere determined approximately for both one percent and five percent deviations from linearityo The variation with time after loading of these linear limit stresses was also examined° Data is presented graphically for all materials. Certain characteristic functions defined by viscoelastic and photoviscoelastic theories are given for Polyester Ernst W. Kiesling Resin CR-39 in both the linear and nonlinear regions. Both these theories are discussed in some detail. Several wave lengths of radiation, covering the visible range, were employed in the investigations of bire- fringence. Two phenomena—-dispersion of birefringence and the possible variation with wave length of linear limit stresses-—could thus be studied. Tapered models were used to obtain birefringence and mechanical data essential to this study. The moire method, used in conjunction with tapered models, provided a con- venient means of obtaining continuous strain-stress data from a single model and loading. NONLINEAR, QUASI—STATIC BEHAVIOR OF SOME PHOTOELASTIC AND MECHANICAL MODEL MATERIALS BY Ernst w. Kiesling A THESIS Submitted to Michigan State University in partial fulfillment of the requirements for the degree of DOCTOR OF PHILOSOPHY Department of Metallurgy, Mechanics and Materials Science 1966 I a flu! ACKNOWLEDGMENTS The advice and encouragement given by the members of the Guidance Committee——Drs. W. A. Bradley (Chairman), G. L. Cloud, J. L. Lubkin, and C. P. Wells——are gratefully acknowledged. Dr. J. T. Pindera gave valuable guidance during the early stages of this research effort. Funds and laboratory facilities for the experimental program were provided by the Department of Metallurgy, Mechanics and Materials Science, and by the Division of Engineering Research. The first year of advanced graduate study was made possible by a National Science Foundation Science Faculty Award. ii TABLE OF CONTENTS Page ACKNOWLEDGMENTS. . . . . . . . . . . . . . . . . . . . ii LIST OF ILLUSTRATIONS. . . . . . . . . . . . . . . . . Vi LIST OF SYMBOLS O O O O O O O O O O O O O O O O O O O O Xi Chapter I. II. III. INTRODUCTION 0 O O O O O O O O O O O O O O O O 1 Objective and Scope Problem of Model Similarity Primary Objectives of Thesis Collateral Objectives Organization of Thesis CALIBRATION TESTS. O O O O O O O Q 0 O O O O O 12 Part A. Description of Material Properties. 12 Introduction Calibration of Mechanical Model Materials The Photoelastic Effect Calibration of Photoelastic Model Materials Complete Description of Material Properties Part B. Calibration Procedure . . . . . . . 35 Possible Calibration Models Model Used in This Study Model Preparation and Clamping Test Procedure—-Photoelastic Calibration Test Procedure-~Mechanical Calibration Materials Tested BIREFRINGENT AND MECHANICAL PROPERTIES OF (ZR-39 o o o o o o o o o o o o o o o o o o 52 Birefringence vs. Stress, Time Constant Birefringence vs. Time, Stress Constant Strain vs. Stress, Time Constant Strain vs. Time, Stress Constant Linear Limit Stress vs. Time iii CONTENTS-~Continued Chapter Page Birefringence vs. Strain, Time Constant Material Coefficients in Linear Range IV. VISCOELASTIC STRESS ANALYSIS . . . . . . . . . 73 Introduction Aspects of the Problem Forms of Representation of Linear Visco- elastic Behavior Forms of Representation of Linear Photo— viscoelastic Behavior Nonlinear Viscoelasticity Nonlinear Photoviscoelasticity Limitations of Results Presented V. VARIATION OF OPTICAL PROPERTIES WITH WAVE LENGTH O O O O O O O O O O O O O O O O O O O l l 5 Investigation Performed Dispersion of Birefringence Variation of Linear Limit Stress with Wave Length VI. DISCUSSION, CONCLUSIONS, AND RECOMMENDATIONS FOR FURTHER RESEARCH . . . . . . . . . . . . 128 Achievement of Objective Experimental Techniques Material Property Descriptions Properties of Particular Materials Recommendations for Further Research LIST OF REFERENCES . . . . . . . . . . . . . . . . . . 145 Appendix I. ANALYSIS OF STRESS DISTRIBUTION IN TAPERED MODEL. 0 O O O O O O O O O O O O O O O O O O 151 II. TECHNIQUES OF MOIRE STRAIN ANALYSIS. . . . . . 161 Difficulties Presented by a "Conventional" Method Suitability of Moire Method Production of Line Arrays—~General Methods Grid Application Method Employed iv CONTENTS—~Continued Appendix Page III. DATA FOR OTHER MATERIALS TESTED. . . . . . . . 176 Polyester Resin PS-l Epoxy Resin PS-2 Polyester Resin Palatal P6-K Figure 2.1 LIST OF ILLUSTRATIONS Representative plot of 5 = 8(6), t = const (tl,t constant load. . . . . . . . . . . . . . 2,t3, . . .) for uniaxial stress, Linear limit stress and relaxation modulus functions of time for uniaxial stress, constant load. . . . . . . . . . . . . . Representative plot of 8 = 8(t), o = const (01,02,03, . . .) for uniaxial stress, constant load. . . . . . . . . . . . . . Isochromatic fringe order n vs. stress for two wave lengths, X1 Variation of linear limit stress with time, based on birefringence . . . . . . . . . Fringe constant f and stress—optic coefficient Co as functions of time for given wave length 0 O O O O O O O O O O O O O O O 0 Order of interference n vs. time for various stress levels, uniaxial stress . . . . . Order of interference n vs. strain for t = CODSt and. A : Al 0 o o o o o o o o o o o o 0 Order of interference n vs. time for 8 = const and )\ = X1 0 o o o o o o o o o o o o o o o 0 Equipment used to scribe scale on models . Sketches illustrating (a) the effect of bend— ing of the clamp jaws, and (b) improvement by modification. 0 O O O O O O O O O O 0 vi and X2. 0 O O O O O O O Page 19 20 21 29 30 31 31 33 33 38 41 Figure 2.12 3.2 3.7a ILLUSTRATIONS--C0ntinued Page Polarisc0pe used in photoelastic studies: (a) light source, (b) condensing lens, (d) dispersing screen, (P) polarizer, (Q1) quarter wave plate, (M) model, (02) quarter wave plate, (A) analyzer, (F) filters, and (C) camera. . . . . . . . . 43 Apparatus used in mechanical model tests: (a) light source, (b) collimating lens, (M) model, (D) "dummy" model or analyzer, (d) condenser lens, and (C) camera . . . . . 48 Optical system used in mechanical model tests. 48 Illustration of method of attaching moire analyzer to model. . . . . . . . . . . . . . SO Isochromatic fringe order n vs. distance along centerline, t = tl,t2, . . . ,tn for CR-390 o o o o o o o o o o o o o o o o o 53 Isochromatic fringe order n vs. stress, t = tl,t2, o o o ,tn for CR_39 o o o o o o o o o 55 Isochromatic fringe order n vs. time, 0 = 01,62, 0 o o ,Gn for CR—39 o o o o o o o o o 57 Moire fringe order m vs. distance along centerline, t = tl,t2, . . . ,tn for CR-39 . 59 Percentage elongation vs. stress, t = tl’t2’ o o o ,tn for CR—39 o o o o o o o o o 61 Percentage elongation vs. time, 0 = 01,02, 0 o 0 ,0n for CR—39 o o o o o o o o o 63 Linear limit stresses vs. time for bire- fringence and strain, time range: 0-240 hours, for CR—39 o o o o o o o o o o o o o o 64 vii ILLUSTRATIONS—~Continued Figure Page 3.7b Linear limit stresses vs. time for bire- fringence and strain, time range: 0-10 hours, for CR—Bg o o o o o o o o o o o o o o 65 3.8 Isochromatic fringe order n vs. percentage elongation, t = tl,t2, . . . ,tn for CR-39 . 69 3.9 The functions CG(t), fO(t), and E(t) for CR—39. . . . . . . . . . . . . . . . . . 72 4.1 Functions 80(t) and D(t) for CR—39 . . . . . . 96 4.2 Plot of (n - no) vs. time for three stress levels for CR—39 . . . . . . . . . . . . . . 107 4.3 Kernels Dl(t), D2(t,t) and D3(t,t,t) for CR-390 o o o o o o o o o o o o o o o o o 109 4.4 Isochromatic fringe order n vs. time, experi- mental and calculated values, 0 = 2,600 pSi, for CR—39 o o o o o o o o o o o o o o o 110 4.5 Isochromatic fringe order n vs. time, experi— mental and calculated values, 0 = 01,62, 0 o 0 ,0n for CR—39 o o o o o o o o o 1].]. 5.1 Isochromatic fringe order n vs. stress, 1 = X1,XZ, o o o ,Xn for CR"39 o o o o o o o o o 120 5.2 Normalized retardation curves for CR—39. . . . 122 5.3 Normalized dispersion of birefringence for CR-390 o o o o o o o o o o o o o o o o o 124 5.4 Isochromatic fringe order n vs. stress, 1 X1,12, . . . ’Xn’ showing linear limit stresses for CR-39 . . . . . . . . . . . . . 126 Al.l Sketch of wedge with concentrated load at the apex O O O O O O O O O O O O O O O O O O O O 152 Figure A1.2 A1.3 A104 A3.1 A302 A3.3 A3.4 A3.5 A3.6 A3.7 A3.8 A3.9 A3.1O ILLUSTRATIONS—-Continued Sketch showing shape and approximate dimension of tapered model used. . . . . . Plot showing variation of the stress con- centration factor k along the centerline of the photoelastic model. . . . . . . . . Photographs of (a) isochromatic fringe pattern and (b) moire fringe pattern . . . Isochromatic fringe order n vs. stress, t = tl,t2, o o o ,tn for PS—lo o o o o o o o o Isochromatic fringe order n vs. time, o = 01,02, 0 o 0 ,6n for PS-l. o o o o o o o 0 Percentage elongation vs. stress, t = tl’tZ’ o o o ,tn for PS—lo o o o o o o o 0 Percentage elongation vs. time, 0 = 01,02, 0 o o ,On for PS—l. o o o o o o o o Isochromatic fringe order n vs. percentage elongation, t = tl’t2’ . . . ,tn for PS—l. Isochromatic fringe order n vs. stress, t = tl,t2, o o o ,tn for PS—2o o o o o o o o o Isochromatic fringe order n vs. time, o = 01,02, 0 o 0 ,0n for PS_20 o o o o o 0 o 0 Percentage elongation vs. stress, t = tl,t2, o o . ,tn for PS-2. o o o o o o o 0 Percentage elongation vs. time, o = 61,02, 0 o 0 ,6n for PS_20 o o o o o o o o Isochromatic fringe order n vs. percentage elongation, t = t t ,tn for PS—2. 2, O O 0 ix 1’ Page 156 158 160 177 178 179 180 181 183 184 185 186 187 Figure A3.11 A3.12 A3.13 A3.14 A3015 ILLUSTRATIONS——Continugg Isochromatic fringe order n vs. stress, t = tl’tz, c o o ,tn for P6‘Ko o o o o o o o n Isochromatic fringe order n vs. time, o = 01,02, 0 o o ,On for P6_Ko a o o o o o n 0 Percentage elongation vs. stress, t = tl’t2, o o I ,tn for p6—K. o o I o a c o 0 Percentage elongation vs. time, o = 01,02, 0 a 0 ,6n for p6-Ko I o o o o o o I Isochromatic fringe order n vs. percentage elongation, t = tl,t2, . . . ,tn for P6—K. Page 189 190 191 192 193 LE LIST OF SYMBOLS uniaxial stress normal stresses along axes of coordinates principal stresses stress tensor or typical component of stress tensor 0 + o 11 * G22 33 5 the stress deviator aka 0. lj ijokk’ uniaxial strain principal strains 8 + 8 11 22 + 5 33 strain tensor or typical component of strain tensor 6 the strain deviator mud Sij ‘ijgkk’ isochromatic fringe order or order of isochromatic or order of inter— ference indices of refraction along coordinate directions moire fringe order time linear limit stress xi n‘ozz)1.o' n [1 Shim if —"'.1‘—-' Small clamping force-f Active model————*- -+———-Dummy model Fig. 2.15——Illustration of method of attaching moire analyzer to model. All tests, both mechanical and optical, were conducted at 68°F : 1°F. With the humidistat existing in the laboratory precise control of the relative humidity could not be maintained. It was observed on a recording hygrometer to vary between 55% and 70%. 51 Materials Tested Because the primary aims of this study are to investigate certain possible sources of error and to suggest ways of improving the accuracy of photoelastic stress analysis, those materials commonly used in photo- elasticity were selected for testing. Polyester Resin CR-39 was the first choice, because of its extensive use in photoelasticity; it is probably the most commonly used of all birefringent materials in this country. This material was supplied by Cast Optics Co., Inc. Also selected for examination were two materials commonly used for photoelastic coatings but believed to be potentially useful also for models. These were supplied by Photolastic, Inc.: Epoxy Resin PS-2 and Polyester Resin PS-l. These, as well as CR—39, are available in sheets of various sizes and thicknesses. Polyester Resin Palatal P6—K, manufactured by Badische Anilin und Soda—Fabrik (Germany), was also tested. The material in liquid form was donated by the manufac— turer. Plates were cast in 1/4-inch and 3/8-inch thick- nesses and then cured. In summary, the materials tested were: Polyester Resin CR-39. Epoxy Resin PS-2. Polyester Resin PS-l. Polyester Resin Palatal P6-K. CHAPTER III BIREFRINGENT AND MECHANICAL PROPERTIES OF CR-39 Birefringence vs. Stress, Time Constant Details of the testing procedure and of the graphical presentation of results were given in the pre— ceding chapter. Discussion in this chapter is therefore limited mostly to the results obtained. The birefringence data presented in this chapter is based on the shortest wave length used in the study—- namely, 410 nm. Since the isochromatic fringe order is inversely related to wave length, the highest possible number of fringes was thus obtained. Figure 3.1 shows the fringe order n as a function of length along the centerline of the specimen. The extensive use of the 410 nm wave length was not anticipated when the test was conducted. Consequently, two pictures with another wave length were taken before the first one with 410 nm at one minute after the load was applied. The accuracy and fineness of the scribed scale (smallest divi— sion 0.20 cm) allowed accurate location of both integral and half—order fringes. The data furnished very smooth curves with no observable irregularities. 52 53 ¢_ .mmumo mom up. . . . .mn.an u n .mcfiaumncmu macaw QUGMDmHol.m> C umpuo mmcfium UHDmEOHLUOmHIIH.m .mam AEoVJ .N. O. m w ¢ JN d d J _ u u d d d d .5 mud n u .ecozfix .8 -mo ..oo ..o ..o /” ‘ ..oo .9 co ///” .0» .oo ... // .0» .oo .1. .8 .oo .13 .oo ..oo. .oogon_ .00 gets do u ‘ upJo obupd 9! w. ON 54 A measuring microscope permitted corresponding accuracy in measuring the width of the specimen at a number of locations. The thickness was measured at several points with a micrometer caliper. The nominal stress was computed from the area and the load of 320.85 pounds (64.17 magni— fied 5.00 times by the load lever). By using the values of the "stress concentration factor" described in Appendix I, the actual stress was computed at each of the points where the area was known. A plot of stress vs. length along the centerline (not shown) yielded no detectible scatter.l The fringe order for each of the times represented in Figure 3.1 was then read at about fifteen values of stress, ranging from 800 psi to 2,900 psi. One additional set of points corresponding to fringe order and stress in the shank is also available. Data was thereby obtained for a plot of fringe order as a function of stress for each time represented. This plot is shown in Figure 3.2. The nonlinearity of the stress—optic behavior is clearly evident in this plot. The linear limit stress at ) one percent deviation from linearity, denoted by n(0££ l 0’ is marked by an "x" on each curve. For the assumption of linearity to be accurate within one percent, it appears necessary that the fringe order remain below seven fringes 1The term "scatter" usually implies the testing of more than one specimen and some sort of data averaging for each point plotted. Tapered models were used in this study to avoid such scatter. Herein, scatter denotes observable irregularities or lack of smoothness. 55 ommlmu “om c#o o o o «mynfl # n p .mmmhwm .m> c Hmouo mmcfiuw UflymeounuomHiim.m .mflm , 2...: b oo~.n 0.8.... comm ooofi 8.6.. com. com oo.¢ I 3.01.5203} .mmumo A: r 1 Aqnzjbflh l-o r. °._Aqubvc llx / I I // r //l ///(I .oo..o no a / I, . ..oo .2 ..o 1 / :00 .008. r /, /.on.oo .1. . I .8 .oo ..3 . /I .oo ..oo. r /I .oo :3. . r b P b P P > _ b IIL [ h p [ b p b p p P L r L t F 2L u ‘upso obugu 0. ON Va 56 within the first hour after loading. This number depends, of course, upon the wave length of radiation and the thickness of the model. It should also be observed that within the first hour after loading a much higher fringe value (or stress) is permissible if a deviation from linearity of, say, five percent is allowed. The linear limit stress at five percent deviation from linearity is marked by an "O" in ) Figure 3.2. The difference between n(OM) and 1.0 n(GM 5.0 is seen to decrease with increasing time. The error in assuming the linearity at small stresses to extend to higher stresses may be considerable. At 3,000 psi, for example, this error is approximately 7% at one minute, 11% at one hour, 28% at 25 hours, and 56% at 240 hours. Birefringence vs. Time, Stress Constant At the eight times represented in Figure 3.2 the fringe order at any stress within the prescribed range may be obtained, thus providing data for plotting fringe order as a function of time with O = constant. Figure 3.3 shows such a plot. Nonlinearity at a given time may again be observed as increasing distance between consecutive curves as the stress level increases. Any irregularities (or "scatter") of raw data would have shown in Figure 3.1 and have been "smoothed” by 57 QQMIINHU HOW no» 0 o o oNbaHb .m> c umpuo mmcflum vameOHSUOmHIIm.m .oflm 332: 06:. n o .mep oom ow. ow. om ov o F P b . . o A u _ u A _ J 4 d 4 d a u - u ‘ q s u q I .5361. .5532 .mm-mo 1 ll.\ :0 00¢ lb ]|I|I|II\ 11 e .803 .b l\\\ T .308. .6 K .uaOON_-b \\\ ... ..uOO¢.-b K m ”m: . .. U 1 2.89.6 w 0 2.002 .6 W O I J ..aooo~..b . 1 N. u _aoo-.b 3.39.9.6 1 1 Jim— i|l\\.. 283.6 .2. 33.6 . 4 r 110m L . . _ _ L P . . _ . p r p _ _ . p . L 58 careful plotting to a large scale (21 in x 33 in). The last two figures were deduced from Figure 3.1; thus no irregularities were apparent in these plots. Strain vs. Stress, Time Constant With the aid of the moire method of strain analysis, evaluation of mechanical properties followed a procedure very similar to that used in evaluating birefringence properties. In Figure 3.4 the moire fringe order is shown as a function of distance along the centerline of the specimen for eight values of time after the load was applied. The grids were apparently distorted in the course of applying them to the specimen and analyzer. This gave rise to an initial moire pattern represented by the lowest curve shown. The first picture of the moire pattern was taken approximately four seconds after the load was applied. With a fringe order of approximately twenty, locating both integral and half-order fringes provided approximately forty points for plotting a curve. This was increased to about seventy-eight points at 150 hours, after which the specimen fractured. The scatter of data points on this plot was again insignificant. Though a more dense moire grid would increase the number of points available for plotting, these additional points would not aid materially in defining the shape of the curve. 59 .mmumu mom a». . . . .mn.ap u p .mcaaumpcmo macaw mUGMMmHo .m> E Hmouo mosauw muaoZII¢.m .mHm A..;oc.. 4 o _ u n ¢ 0 m s m b p p p . p . P 0 _ O a . q a 4 a a q 4 d a . _ q . 4 TI IlIIII-I‘ull‘liuil‘ll‘ll‘lu‘I-II' L 1' . I \‘\‘x I] V\ a. II v I l vl Itl m H 65):: 08 ....N. v mml mo 1 T 1 1 im— /////Ioo .co .0050 H ..O to HION /////l. .oo.o.go 1 / :00 OO: _ 11 VN /|l.. ...OO 00 no. 1 .oo :00 lrmN L 1. . .oo ‘00— L .oogon_ ..mn r L .l .1 h _ _ p p _ b b L p L . .r a b . 0” m ‘Joplo obugu ”you 60 Stress was plotted as a function of distance along the centerline of the specimen. The strain at a given stress and time is represented by the slope of one of the curves at a particular point. Slopes were measured by a graphical technique, employing a set of mirrors mounted on a 90° machining block. One mirror was placed tangent to the line at the chosen point, the other perpendicular to it, thus yielding an image of the line tangent at the point and an image continuing the line at the point. This method of measuring slope is, of course, susceptible to significant error. However, readings could be repeated to within about four percent. The possible increase in accuracy resulting from a sophisticated numeri- cal evaluation of the slopes did not seem to be justified. The slope evaluations yield data for plotting strain (or percentage elongation) as a function of stress for each of the times represented in Figure 3.4. Such a plot is shown in Figure 3.5 with the data points indicated. Some lack of smoothness is seen to exist. Because of the ini— tial moire pattern, the data points actually represent the difference of two slope measurements, increasing the possible error. Despite this the scatter does not seem excessive, and the results are believed quite reliable. Nonlinearity of the mechanical behavior at high stresses is again rather obvious. The departure from 61 OOOn T - T OONN oovw . ooa mnsmo .mmnmu now up. . . . .Nu.Hu u p .wmmuym .m> coaummsoam manpcmuummllm.m .OHm com. :3. b 009 CON— o.oaquvw II o 33.3w I u x rO ....MO :66 .10.. ..N uouobuma obmuooud 62 linearity (vertical deviation from extended straight line) seems more rapid than for birefringence, as indicated by comparison with Figure 3.2. The stress at one percent deviation from linearity 8( ) is again marked by an CM. 1.0 "x" on each curve and 8(OM)5 0 by an "0." Linear limit stresses will be further discussed in a subsequent section. Strain vs. Time, Stress Constant From Figure 3.5 data can be deduced for a plot of strain (or percentage elongation) vs. time for a fixed stress level. Figure 3.6 shows such a plot for twelve levels of stress. Deviations from linearity again appear as increasing distances between consecutive curves at higher levels of stress. Since the plot of Figure 3.5 "smooths" the measured strain data, the data points of Figure 3.6 furnish regular, well—defined curves. Linear Limit Stress vs. Time After the optical and mechanical linear limit stresses are determined from Figures 3.2 and 3.5, respectively, it is possible to show them as functions of time. This is done in Figures 3.7a and 3.7b; different time scales are necessary to show clearly their variation throughout the time interval for which they were determined. 63 oQMI-NHU .HOM CD» 0 o 0 «anHD H .0 .mEHO .m> coarmmcoam mmmrcmuummllw.m .mam :32: 25... . o: 8. oo. 2. on 0.. om. o _ — p P I p _ _ - o u . a q q d _ u u _ _ u _ _ Only—0 «.0 .3 com .6 .mn 000 «.0 ¢.O .3000. «b .3 cow. .6 mo .3 cow. .6 .3 cow. .6 0.0 .3 ooo. . b l.l||\.|ll..||l\‘ .3 ooom ..b 0.. .3 com». a b N._ .3 83 .b . 1V.— .3 003 ..b , -9. .0.— .3 anew . b T ..O.N p n u b n n n — p p b h p uogwbuma obmuoomd 64 .mmimu How .mhson owmlo "mmcmu mEHp .sflmupm paw mucmmcfiummufln mom mEHb .m> mmmmmupm pHEHH uanHQIlm>.m .mflm Amazon. mEHH cam oom 06H oma om as o o u q a d u d u q 4| d — - u (H -ms. -o.H .m.H u .o.m ea ofla n K .mmumo 4w y.m.m 65 .mmlmu mom .munon OHIO "008mm wear .Qflmuwm 0cm mucmmnflummufln How mEHp .m> mmmmmuwm uHEHH Hmmcwallgb.m .mflm AmmuScHEV mEHB com com 00¢ com com OOH as 0H6 n K .mmumo _ A] . _ _ q q d _ _ q q — _0I (Isd) 66 The linear limit stress at one percent deviation from linearity is seen to be considerably lower for bire- fringence than for strain. Though little information is available for comparison, such a wide difference was not expected. Coolidge [7] reported that for a given rate of loading the stress-strain curve for CR-39 is a straight line up to about 3,000 psi, whereas linearity of the stress-fringe relation extends to "somewhere between 2,500 and 3,000 psi." The resin he tested was manufactured by the Pittsburgh Plate Glass Company. The method of testing employed in his investigation would not seem to permit adequate allowance for creep effects. Frocht [18] found that in Bakelite (BT—6l-893) the "proportional limit" for birefringence is over twenty percent higher than for strain. Creep of this Bakelite was apparently considered negligible. Approximate values of n(oaa)l.0 for CR-39 manufactured in Great Britain may be deduced from data presented by Pindera [47]. The values seem to be consider- ably higher than those of the American—made CR—39 tested in this study. However, the ultimate strength is also considerably higher for the British-made resin. It must be concluded that the linear range of behavior, as well as other properties of CR—39, varies considerably with different manufacturers. Properties may vary from one lot to the next with the same manufacturer. Age and storage conditions may also affect some or all of 67 the properties. Pindera's CR-39 is known to have been stored for five years, whereas the material used in this study was tested within six months after purchase. The qualitative results of this study seem consistent with earlier findings; the quantitative differences emphasize the need for careful calibration. Comparison of Figures 3.2 and 3.5 seems to indicate a more gradual departure from linearity in the birefringence—stress relation than in the strain-stress relation, especially at times greater than one hour. The linear limit stresses for birefringence and strain at five percent deviation from linearity are more nearly equal than those at one percent, except at very short times. This can be seen in Figures 3.7a and 3.7b. The crossing of the two curves at five percent deviation is puzzling and may be due to accumulated experimental error. Two quantities are being compared which are difficult to determine in the first place and which are of roughly the same magnitude. Irregularities are therefore not surprising. The crossing of the curves may actually portray the time behavior of the material. This does not seem entirely implausible but may be worthy of further detailed study. Birefringence vs. Strain, Time Constant From the plots previously shown, the relation of birefringence vs. strain at a given time after loading may 68 be deduced by the method outlined in Chapter II. Such a plot is shown in Figure 3.8. The largest strain encountered at one minute after loading was slightly greater than one percent. The relations are, of course, linear up to the fringe values or strains corresponding to the optical linear limit stress. Then there is a "breaking over" toward higher strain, indicating that at a given time after loading the strain increases faster than fringe value with increasing stress. Thus birefringence deviates more gradually than strain from linearity. No reference to such behavior could be found in the literature for CR-39, either to confirm or contradict these results. The inflection of the curves at still higher values of strain (or birefringence) may warrant additional confirmation. It is not believed attributable to experi- mental error. An error analysis would be somewhat diffi— cult to perform and would be of questionable accuracy. A statistical analysis of the scatter of the data points on the birefringence—stress and strain-stress plots might provide a meaningful estimate of the consistent error committed. Because of the regularity of the birefringence— stress plot, such analysis should seemingly be applied only to the strain—stress plot. But errors in measuring the slope of the moire fringe order vs. length curves are mostly reSponsible for the scatter of the strain—stress plot. The repeatability of the slope measurements (within omM'mU “om Spa 0 o o oN“aH# " p «GOHpmmGOHm mmmpcmuumm .m> Q umouo mmCHum UHmeOMQUOmHllw.m .mflm cozoocofi 32:33.". 69 tN m; m; «4 may 05. .00 . F ~ ~ E. _ b 0 d J a a u — d J _ — A q u — H u _ _ _ , .\ i, \x I oogon. .. . \.\\.\ .. . 1 oo oo. .\ 1 . ... II 4 . I ..oo.oo ... i . \._\\.\ .-m I oo _ o i . .. . l : . z \\ I . 11m 1 #3 r 1 r --.... I iwm_ I .536... £5.03... .mmumo - p b P _ _ _ _ _ p _ _ _ b _ _ P — h _ _N u ‘JOpJO abupd 70 about 4%) offers some estimate of the possible errors in strain values. Consistent errors of i 4% could cause the inflection, but errors of this magnitude are not believed present. Identical techniques were employed in testing the other materials (see Appendix III); no inflection appeared in birefringence-strain plots. Thus if consistent error is responsible for the inflection of the plot for CR-39, such error is apparently unique to this one test. Further reduction of data, such as that necessary for a plot of fringe value vs. time for constant strain, does not seem advisable. Data for such a relation can be obtained directly from a relaxation test, and the plot of birefringence vs. strain for t = constant (as shown in Figure 3.8) can be obtained in a single step. Relaxation tests are thus recommended to determine whether the inflec— tion indicated in Figure 3.8 is a true representation of this material"s behavior. At a given strain Figure 3.8 shows a decrease in birefringence with increasing time. This is expected for CRm39 by comparison with results presented by Clark [3] but, as he points out, is not true of all materials. Material Coefficients in Linear Range The photoelastic and mechanical material coefficients, such as E(t), fo(t)’ and Co(t)’ can be deduced from the curves presented. As indicated, these 71 are functions of time, rather than constants as is sometimes assumed. It should further be emphasized that they are defined for the linear range of behavior. From the slopes of the linear portions of the curves in Figures 3.2 and 3.5, fO yielding CO and E(t) are deduced. These functions are plotted in Figure 3.9. Their strong time-dependence during the first two or three hours, the time during which most static tests are performed, should be noted. The possible error which may result from assuming these functions to be constants, independent of time, is evident. 72 mmlmu MOM. ELM 9.8 n b Ar. 0 .Apvpo mcowpussw wraqam.m .mwm AmmDSCHE. meH 600.0H ooo.a co. OH o.H fl1__-_ a __.u_._ a fi —___q|¢_ _ _ ~_fld_41d _ 4:4... _ H a an mm.o n 6 .56 ans n . .mmsmo .-.. - /hmmvmuoH . .nvm Hmouo x -Ca. R . 6 . . , A Q. .muon . .3. u 7 . H! . <\\.\\ . 4 g . . . HQHO on SA. .. \\ a\\\fl A U J4 .VVIQH . vabw.l .xi .. . / —.PF-.. - _..... _ L . _CLP._-LI'|_IIIII'1._C|.C_ _ _ ___._.p h CHAPTER IV VISCOELASTIC STRESS ANALYSIS Introduction From the discussion and the graphical presentation of results in the two preceding chapters, the change with time of the mechanical and optical properties of the mate- rials is quite apparent. This viscoelastic behavior is manifested as creep under constant stress or relaxation under constant strain. In most discussions, strain and stress are considered the time—dependent functions in creep and relaxation tests, respectively. The simplest way to describe the viscoelastic behavior is then to define the extension modulus D(t) or one of the shear or bulk compli— ances or moduli (to be defined) over the whole range of time and temperature. However, other time—dependent func- tions such as birefringence may be considered, with an analogous photoviscoelastic description given in terms of the stress-optic or strain-optic coefficients in creep or relaxation. In the following discussion of viscoelastic- ity, strain is usually considered the basic time—dependent function in the discussion of the creep tests performed in this study. Birefringence data was also obtained, and the isochromatic fringe order is considered the basic 73 74 time-dependent function in a similar discussion of photoviscoelasticity. In this chapter some elements of the existing theories of linear viscoelasticity and photoviscoelasticity are discussed. Graphs of some "characteristic functions" defined by these theories are presented. Nonlinear theories are then discussed and applied to the optical creep test of CR—39. Though viscoelasticity and photoviscoelasticity have much in common, discussions of them are separated in the following. The similarities and differences are pointed out where they do not seem obvious. Aspects of the Problem The problem of viscoelastic stress analysis has three aspects [ll]: determination of material properties, interrelations among the various descriptions of visco» elastic behavior, and the methods of stress analysis. The first of these, determination of material properties, is the main concern in this study. The other two aspects warrant consideration even in a study of this type, since they determine the form of representation of material prop— erties most suitable for a particular problem, as well as define the range of time, stress, temperature, etc. for which material properties must be determined. The first two aspects might indeed be considered as one [27]—-namely, that of specifying the constitutive equations of the 75 material which are necessary for the analytical solution of boundary-value problems. Most of the theoretical development of viscoelastic— ity,and surely most of the problems which have been solved, assume linear viscoelastic behavior--i.e., the time- dependent functions are approximately linearly proportional to the applied constant stresses or strains.l Many photo- elastic and model materials, including those tested in this study, are found to be linearly photoviscoelastic or visco- elastic, provided the applied quantities do not exceed certain limiting values. The theory of linear viscoelas- ticity can then be applied to solution of boundary-value problems involving such materials. Although there does not exist as yet a systematic or unified body of theoretical results on linear viscoelastic- ity comparable to that available in classical elasticity theory, it may be considered well-established. Work such as that of Gurtin and Sternberg [27] established a firm mathematical basis for the theory and gave a clear formula- tion of the problem. 1Since the linear limit stress (in a creep test) decreases with time, the designation "linearly viscoelas- tic" is more restrictive than "momentarily linear"--i.e., the former implies specification of a time interval throughout which behavior is linear. Within this time interval the behavior will be momentarily linear for a wider range of stress than that for which it is linearly viscoelastic. Reference to Figure 2.1, page 19, might clarify this assertion. 76 The isothermal stress analysis problem for an isotropic medium is formulated as follows: Given a body, occupying a closed region of space in its undeformed state, subjected to prescribed surface tractions Ti(xk’t) over part ST of the boundary and to prescribed surface displace- ments Ui(xk,t) over the remaining part Su of the boundary, determine the stress and strain distributions throughout the body. If the body is loaded or displaced at t = 0, then initial and boundary conditions take the form: 0.. = 8.. = u. = 0 for t 55 0 (4.1) ij ij i oij(xk,t)nj(xk) = Ti(xk,t) on ST for t :> 0 (4.2) ui(xk,t) = Ui(xk,t) on Su for t >1 0, (4.3) where ui = displacement component in the direction of the ith coordinate (i = 1,2,3), nj = unit outward normal vector on the boundary, xk = kth coordinate of point. The equations of motion and the (small) strain—displacement equations are: ...... x, = (4.4) ngj + l P <§ 2 011. d)u. =é l+—"‘l . (4.5) 13 (jx. (5xi where Xi = body force component in the direction of the 1th coordinate, f) = density of medium. Finally, the constitutive equations of the viscoelastic material may be stated in general as: P(sij) = Q(eij) (4.6a) P'(Oii) = Q'(8ii), (4.6b) where _ l . sij — Oij --§ 6iijk’ the stress dev1ator, _ l . . eij — Sij --§ 6ijgkk’ the strain deViator, P,P', 0,0' = linear operators with respect to time (to be further explained). The solution of the stress analysis problem consists of solving the system of Equations (4.2)-(4.6). Time vari- ations, if any, of the traction and displacement boundaries must be prescribed. This can cause severe complications; in most solved problems such variations are not considered. In quasi-static problems the inertia terms in the equation of motion are negligible; the right-hand side of Equation (4.4) then becomes zero. Most problems so far solved are of this type. 78 In the methods of stress analysis mentioned at the beginning of this section—~i.e., in arriving at a solution of Equations (4.2)-(4.6)--severa1 possible approaches have emerged with, of course, a number of variations of each. In one such approach the governing equations are solved directly, which is sometimes possible by integrating sepa- rately with reSpect to the time and space variables. Lee [33] gavra an example of this in an excellent survey article. A more common approach applies the Laplace trans- form to the system of equations, thus removing the time- dependence, so that the viscoelastic problem becomes an elastic one in the transformed variables. The so—called "correspondence principle" is thereby invoked. This method applies directly only to problems in which the portions ST and Su do not vary during the loading process [34]. A dis- cussion of this approach, along with reference to a number of solved problems, may again be found in the survey article by Lee [33]. In addition to the mathematical complexities often encountered in the two methods mentioned above, they require the expression of material behavior in mathematical form-~i.e., the viscoelastic operators of Equations (4.6a) and (4.6b) must be known. Such expressions can easily be too complicated to handle in a mathematical solution or can be too simple to adequately describe the behavior of the material, or both. This is particularly true in dynamics 79 problems where very short-time response is required. Physically meaningful solutions, nevertheless, have been obtained either by fitting mathematical functions to experimental data on material properties or by using such data directly in numerical solutions with the aid of a computer [34]. This latter method removes the limitation to the type of boundary—value problem which can be trans» formed into an associated elastic one. Another approach is the completely experimental one. The importance of such an approach becomes apparent when the complexity of the viscoelastic stress analysis problem is considered. As in elasticity, solutions of only those problems with simple boundary conditions and geometry seem feasible analytically. When the correspondence principle is used, the viscoelastic solution indeed depends upon the availability of the solution to the associated elastic problem. It appears then that the problem of viscoelastic stress analysis is generally more complex than, or at best just as complex as, the elastic one. One experimental approach for viscoelastic stress analysis is photoviscoelasticity-—an extension of photo- elasticity for elastic analyses--a logical extension indeed in View of the fact that most photoelastic model materials are inherently viscoelastic. Some aspects of this experi- mental approach are given in the following sections. 80 Stress, strain, and birefringence in a viscoelastic material are three uniquely interrelated second-order tensors. Mindlin [39] derived a stress-strain-optic law in the form of second-order linear differential operators for an idealized four-element model, assuming the material to be incompressible. Mindlin's work and that of others who followed are reviewed by Daniel [11] and extended to forms suitable for stress analysis purposes. Daniel gave the stress-optic law in both differential and integral form. The experimental approach necessary for calibration of the material and solution of the stress analysis problem was outlined and then applied to a dynamics problem. The following sections give a rudimentary account of some methods of representation of the viscoelastic and photoviscoelastic properties of materials. Only a few of the many possible representations appear, and these are specialized to the needs of this study—~namely, to the iso- thermal, quasi~static case. Reference is made to a number of sources where more general developments are given. Forms of Representation of Linear Viscoelastic Behavior Among the most comprehensive works in linear visco- elasticity are those by Leaderman [31], Alfrey [1], Gross [25], Staverman and Schwarzl [62], Ferry [1?], and the three-volume one edited by Eirich [16]. Reference to many recent publications may also be found in Lee's paper [33]. 81 Equations (4.6) gave the constitutive equations for a viscoelastic material in terms of deviatoric and dilata— tional parts of stress and strain. The corresponding equations for an isotropic elastic material are: 5.. = 2Ge.. (4.7a) lJ 1] 0.. = 3K8.., (4.7b) ii ii where G = shear modulus, K = bulk modulus. Equations (4.6) may conveniently be written: P(sij) = 20(eij) (4.8a) P'(oii) = BQ'(sii), (4.8b) where P, P', Q, and Q' are again linear operators with respect to time, identical to those of Equations (4.6) except for the constant coefficients. Hereafter any refer- ence to these operators assumes those of Equations (4.8). In differential—operator form these operators are expressed as: m r n r P = E p & , Q = Z q & r tr 1: C] tr r=0 9 r=0 , (4.9a,b) 82 m' n' r r P' = E p} ‘3 r’ 0' = E q; (3 r’ (4.9c,d) r=0 (3t r=0 6t where the coefficients pr, q pg, and q; are constants of r’ the material. The number of such constants-~i.e., the values of m, n, m', and n'--necessary to describe a mate— rial depends upon the viscoelastic behavior of the material and, of course, upon the accuracy desired. Frequently, the number required is high enough to make the mathematical treatment of the equations formidable. The constants in the differential operators can be related to an idealized mechanical system composed of springs and dashpots. These mechanical systems then serve as pictorial representations of the material behavior. For such purposes the constants are usually specialized to correspond to the elements of simple mechanical systems, such as the Kelvin model (spring and dashpot in parallel), Maxwell model (spring and dashpot in series), or various combinations of these. The idea of the representation by mechanical models can be extended to an arbitrarily large number, indeed to an infinite number of elements. The operators are then characterized not by a finite number of constants but by continuous functions called distribution functions [62]-—"distribution of retardation times" or "retardation spectrum" for the continuous Kelvin model used to describe creep, and "distribution of relaxation times" 83 or "relaxation spectrum" for the continuous Maxwell model used to describe relaxation. The spectra give a picture of the springs and dash— pots of the model which in turn can be thought of as representing molecular processes. Hence the spectra give a close representation of the molecular processes occurring in viscoelasticity, a fact which is important to the chemist or physicist relating chemical constitution of a material to its properties. Perhaps this is the best justification for their use to describe material behavior. A more general representation is given by visco- elastic operators in integral form, which result from considering Boltzmann's superposition principle [1] for continuous stress or strain histories. Such a super- position principle is valid for (or equivalently describes) linear viscoelastic behavior. In integral-operator form, the deviatoric strain response to any time-varying stress sififit) is given by: '.) t .. l dsij(T) Slj(t) :; 'Z J(t-T) T dT, (4.10) ...CD where J(t), called the "creep compliance" in shear, repre- sents the strain response to a unit sustained shear stress. A similar expression for pure dilatation is: 84 1 t ddii(T) 8ii(.t) = '3' E(t-T) —-a-:E--—— CIT, (4.11) where E(t) is the "bulk compliance." When the material has no stress or strain history-- i.e., it is initially "dead"-—then the lower limits of the integrals are replaced by zero. Corresponding equations may be written in terms of "relaxation moduli" for the stress response to time—varying strain. Such moduli, as well as the compliances defined above, usually characterize the longutime behavior of materials. Though they could be used to describe dynamic behavior as well, provided their time variation for suffi- ciently short times were known, they are usually abandoned in dynamic studies in favor of complex moduli and compliances. These may be obtained from the response to sinusoidal loading of variable frequency. Since only creep tests were performed in the present study, only the creep compliances are of value here. It should be mentioned that all of the moduli and compliances of an isotropic, iscoelastic medium are related and can, in principle, be found when only one of these so~called [62] "characteristic functions" is known for all times (or frequencies). Gross [25] presented many transformation formulas for this purpose. Not only are some of the transformations difficult to carry out, but 85 usually no one characteristic function is known for a sufficiently wide range of time or frequency to make all the transformations practicable. Some functions might then be thought of as complementing others for different time intervals rather than being equivalent. The Equations (4.10) and (4.11) for deviatoric and dilatational strain response to time-varying stress may further be specialized for the case of uniaxial tension. An "extension compliance" D(t) can be deduced from the creep and bulk compliances, since a deviatoric and hydro» static stress can be superposed to give uniaxial tension. In integralwoperator form, the axial strain response to timenvarying uniaxial stress is: t s(t) =f D(t-T) Egéll d’L'. (4.12) 0 Note that the lower limit of zero has been employed, assuming no previous stress history. [If the loading consists of the sudden application of stress 00 at time zero which is then maintained constant, the strain response found from Equation (4.12) is: E(t) = GOD(t). (4.13) The extension compliance will be independent of 0 within 0 the range of stress for which the material is linearly viscoelastic. 86 For relaxation-type tests the stress response to time-varying axial strain is: t - ds(1) 0(t) -;/h E(t-T) —7fi7—— d1, (4.14) 0 where E(t) is the relaxation modulus or extension modulus. For a suddenly applied (at time zero) strain which is then maintained constant the stress response is given by: C(t) = 80E(t). (4.15) The extension compliance and modulus are formally t ‘jf E(t-T)D(T)dT = t. (4.16) For moderate rates of creep an approximate relation is [11]: E(t) ’5: W0 (4017) To Show how the creep and bulk compliances can be deduced from the extension compliance (and other easily accessible information) for the special case of sudden application of constant load at time zero, it is necessary simply to insert the deviatoric and dilatational parts of 87 the stress and strain tensors for uniaxial tension into the appropriate operators (4.10) and (4.11). The tensors are: 60 O 0 811 0 0 0 O 0 O 0 s 33 Now from Equation (4.10) the (deviatoric) shear creep compliance is (see also List of Symbols): J(t) - 2eij(t) _ 2ell(t) _ 2811(t) - 822(t) — 833(t). ‘ - " W " _____._._ _ 3 5ij 511 G0 (4.19) and the bulk compliance from Equation (4.11) is: 3s..(t) 3[s (t) + s (t) + s (t)] E(t) : “ii.— : ll 22 33 o (4020) O'.. 0 ii 0 Thus if one of the transverse strains is measured along with the axial strain, then the other transverse strain can be computed (assuming isotropy); or equivalently, Poisson‘s ratio1 n(t) may be obtained. When Poisson‘s ratio is employed, the above equations become: 2811(t)[1 + u(t)] J(t) = (4.21) O' O l o O C O o "POisson's ratio" is generalized beyond ltS normal meaning and is here defined simply as the negative ratio of the lateral to the axial strain. For other definitions see page 73 of Stuart [63]. 88 and 3811(t)[l - 2u(t)] E(t) = O . (4.22) O For many polymers Poisson's ratio is found to be nearly constant with time. Unplasticized high polymers at room temperature are in their glassy state [64]. It is known [62] that Poisson's ratio may quite accurately be assumed constant over a wide range of time for such materials. Daniel [10] found it to vary little over four decades of time in a lowmmodulus material with true axial strains as high as fifteen percent. Equations (4.21) and (4.22) are even further simplified if the time-dependence of PoissonVS ratio is neglected. A simple tension test, yielding the extension compliance, along with the known or measured value of Poisson“s ratio, then suffices to com- pletely characterize the linear viscoelastic, mechanical behavior of an isotropic material. Before proceeding to the forms of representation of photoviscoelastic properties, one further point regarding the mechanical behavior should be noted. In the most general case of loading, the principal axes of stress do not coincide with the principal axes of strain in a visco~ elastic medium. This could be expected from a comparison with plasticity theory, where such coincidence breaks down unless the directions of principal stresses are constant. Although few examples of this have been observed in 89 plastics, Frocht and Thomson [23] found such a lack of coincidence in cellulose nitrate when the loading was such that the directions of principal stress rotated. Both stress and strain enter only one set of equations in the formulation of a viscoelasticity problem [Equations (4.l)—(4.6)]-—namely, the constitutive equations. There the stresses and strains are divided into deviatoric and dilatational parts. Hence it would seem, in principle, possible to solve analytically the general viscoelaSoic problem wherein principal axes of stress and strain do not coincide. The method used to solve the equa- tions, however carefully chosen, would likely lead to a lengthy and involved analysis. Experimental methods, one of which is briefly discussed in the following section, offer an alternative approach. Even here, a number of complications arise. On the other hand, Mindlin [39] showed that if the stress can be represented as a product of two functions, one of which depends only on position coordinates (xk) and the other only on time, i.e., oij(xk,.) = Cij(xk)f(t); (4.23) then the principal axes of stress and of strain coincide. This condition excludes, for example, moving loads, which would cause principal stress directions to change, and 90 initial stresses, unless these stresses are of the same type as those represented in Equation (4.23)o Forms of Representation of Linear PhotoviECOelastic BehaviOr As stated previously, stress, strain, and bire» fringence are three uniquely interrelated second-order tensors. Daniel's representation [11] in integral form of the stress~optic law for viscoelastic materials seems most practical for stress analysis purposes. Though his work draws upon that of Mindlin [39] and others, his formulation seems to be the best suited for inclusion in this study. A slightly different notation is used in the equations pre- sented below, reflecting that of Theocaris [64]. If the direction of light propagation is considered the zwdirection of a cartesian system of axes, then the “x I’D 1.- Q! (‘1' A Wire retardation (birefringence) observed in the usual applications of tw0wdimensional photoelasticity is the maximum difference of birefringence in two mutually perpenm dicu lair di' 6 m fictions in the xmy plane. These directions are Gal. ed the secondary principal optical directions, and the I 3 ... corresponding values of the index of refraction are called the secondary principal values. Similar terminology is used for stresses and strains. Stresswbirefringence relations may now be written 91 t . _ d OXX ..., oyy —f BUR-1:) a? (nxx — nyy) d1. (4.24) 0 where O 3 o = normal stresses along axes of coordinates, xx yy n , n = indices of refraction along coordinate XX YY directions, Bo(t) = time—varying difference in secondary principal stresses for unit sustained birefringence; or as: t d Oxx - ny =‘jf Bo(t‘1)'3? [nO(T) cos 2$n(T)] d1, 0 (4.25) where V U nO(t) (n - n difference in secondary principal values of index of refraction (relative birefringence), @n(t) 2 angle between secondary principal optical directions and coordinate axes. The secondary principal stress difference (01 — O2) may be related directly to the relative birefringence by: 6 92 t 2 Cl - 02 = j[. BG(t—T)-%? [nG(T) sin 2¢n(T)] d1 0 l t 2 .2 + ~/F BO(t—T)-%? [nO(T) cos 2@n(T) dT] . (4.26) O The principal stress directions are available from: t BG(t-T)-%? [nG(T) sin 2®n(1)] d1 5% tan 2®G(t) = t a BO(t—T)-§? [nO(T) cos 2¢n(T)] dT (4.27) where @O(t) = angle between secondary principal stress directions and coordinate axes. Finally, the shear stress is given by: t OXY =-%‘jf BO(t-T) g? [nO sin 2¢n(T)] dT. (4.28) 0 Thus the photoviscoelastic determination of the secondary stress difference at a point in a viscoelastic body, as a function of time, requires a continuous record of birefringence and isoclinic data in addition to 93 knowledge of the stress—optic coefficient B0(t). This is in contrast to photoelasticity where such a stress differ— ence is determined by the momentary values of fringe order and Cc(t). To separate the principal stresses the shear difference or oblique incidence methods commonly used in photoelasticity can be applied [21]. Some modifications accounting for the difference of @n and $0 may be necessary. This accomplished, the strain can be determined using the theory already presented. An alternative proce- dure for determining strain is to measure it directly, possibly by employing the moire method. Such an approach would require no prior knowledge of mechanical properties. Evaluation of Bo(t), the time-varying difference in principal stresses for unit sustained birefringence, in principle, could be determined directly from an optical relaxation test. But such a test, wherein birefringence is maintained constant, is very difficult to conduct. However, if a(t) represents the time-varying birefringence for a unit sustained difference in secondary principal stresses (creep test), then d(t) and 80(t) are related as follows: t [ or.(t-1:)BG(T)d1: = t (4.29a) O or 94 t _jr a(T)BO(t-T)dT = t. (4.29b) O For moderate rates of creep, [30m Egg, (4.30) which permits the determination of Bo(t) by means of a relatively easy-to—perform creep test such as the one performed in this study. For a body with symmetrical shape and loading ¢n(t) can be made zero and Equation (4.26) becomes: t d 01 - 02 =-]F BG(t-T)-a? [nd(r)] dr. (4.31) O For a suddenly attained birefringence nO which is main— tained constant from time zero (optical relaxation test) Equation (4.31) becomes:1 01 — 02 = Bo(t)n0. (4.32) The analagous equation for an optical creep test is: n(t) = a(t)(dl - 02). (4.33) 1Comparison with Equation (2.8) reveals that 2fq(t) E(t) = —-a———. 95 In the present study a uniaxial stress state was employed and a(t) was determined from n(t) = a(t)O, (4.34) which, by use of Equation (4.30), determined Bd(t). Plots of Bo(t) and D(t)l as functions of time are shown in Figure 4.1. Since these characteristic functions are defined for linear optical and mechanical behavior, they can describe the behavior only for a limited range of stress and time as determined in the previous chapter. Some similarities and differences in the treatment of linear viscoelasticity and linear photoviscoelasticity may be noted. It has been pointed out that, generally, mechanical coincidence in a viscoelastic body does not prevai1--i.e., the principal—stress and principal-strain axes do not coincide. This is in contrast to elasticity where they do coincide, at least for isotropic media. Furthermore, in photoelasticity the principal—optical axes coincide with the principal-stress axes so that the meaning of isoclinics is clear. But the question now arises whether the breakdown in mechanical coincidence is accom- panied by a breakdown in optical coincidence; or equiva— lently, what is the meaning of the isoclinic parameters in the presence of general viscoelastic or plastic behavior. Do the isoclinic parameters give the directions of the Y— 1Refer to Equation (4.13) for definition of D(t). 96 .mmumo now ““00 0am Anvom chHpoczmaln.a .mflm Asses @549 000.04 000.4 00H 04 0.4 H.0 .14___ a _ __.4__‘ _ _ __.___—fi_ _~_.___ d _ ~_____q _ I an mm.0 n a 1 es 0H4 u 4 0mnmo x umpuo x NCH 0 1 an 0:04 . inc 0 as m 04 . ““00 0 Eb___ _ - _ .o Auvo «uvm —-___——- _ _____._ p p c . “#000 u Assn __r____ _ l 97 secondary principal stresses, the secondary principal strains, or neither? The theory above assumes neither. Equations for the strain—optic laws similar to the stress—optic ones above would require the introduction of another angle $8, the angle between the secondary principal-strain axes and the coordinate axes. The above question, rephrased, becomes: is Tn = m or is Qn = T G, 8, or is $0 f @n # $8? In the most general case it must be assumed that the last condition prevails. Theocaris [64] pointed out, however, that the "glassy" region of visco- elastic behavior, where deformation is predominantly elastic, and the "rubbery" region, where the creep rate is quite low, correspond to quasi-elastic conditions wherein the axes of secondary principal stress, strain, and bire- fringence coincide. He observed marked deviations from coincidence in the "transition" range, where creep rate is high, in a highly (60%) plasticized epoxy polymer. Frocht and Thomson [22] found similar deviations in an American cellulose nitrate but found that after several hours the secondary principal optical directions coincided with secondary principal stress directions--i.e., @n = $0. But in a German cellulose nitrate Frocht and Cheng [19,20] noted no such time delay. Although mechanical coincidence broke down when principal stresses underwent rotation, optical coincidence (@n = $0) continued to exist. _Such .results are quite remarkable and far—reaching in regard to 98 the applications of photoviscoelasticity or photo- plasticity. It appears that time spent in material selection might be profitably invested. Nonlinear Viscoelasticity The theories of linear viscoelasticity and photo— viscoelasticity, which were briefly introduced in preceding sections, describe the mechanical and optical behavior for only limited ranges of stress and time. Even within these ranges some anomalies are probably present but cannot easily be detected. Thus the descriptions "linear" and "nonlinear" are somewhat arbitrary. The nonlinear region might be defined as that region where deviations from linearity can definitely be detected or cannot be over— looked. That such a region exists for some of the mate— rials tested in this study becomes quite evident from inspection of the plots of Chapter III or Appendix III. Though considerable effort has gone into formulating theories of nonlinear viscoelasticity, the subject is not yet well-developed nor is there agreement on the most satisfactory approach. This is to be expected in view of the relatively recent beginning of any concentrated research efforts, the many types and causes of deviations from linearity, and the mathematical complexity of the subject relative to its linear counterpart. One factor limiting development is the lack of accurate experimental data. As late as 1965 Theocaris [64] 99 wrote, concerning nonlinear behavior of polymers in their glassy state where strains are relatively small: "Experi— mental evidence is rather sparse and the only significant contributions which exist concern the nonlinear behavior in the rubbery zone." This study hopefully provides some evidence of the type mentioned. Most attempts to describe nonlinear behavior fall into one of the three categories: (1) empirical power laws (see Marin and co-workers [38]), (2) postulated behavior in terms of models (with nonlinear springs and modified dash- pots), and (3) integral representation in the form of a modified superposition principle. All three of these have their roots in corresponding representations of linear behavior. Of the three, the last one seems most natural and most successful; further discussion will be limited to it. Green, Rivlin, and Spencer [24] established the basis for the nonlinear theory in a series of highly mathe— matical articles. Quoting Theocaris again, ". . . it seems that only simple cases may be satisfactorily considered by this theory." The theory has been successfully applied by a number of investigators to such simple cases. Ward and Onat [69] presented a clear and concise summary of the theory as it applies to the behavior under simple stress conditions. Some elements of their presentation are reflected in the one below. Lockett [35] considered a more 100 general case and showed how the material functions (properties) for a general triaxial loading can be deter- mined experimentally for a material whose response can be adequately expressed by a constitutive equation involving integrals of the first, second, and third orders. The constitutive equation is expressed in matrix form; the number of tests shown necessary for completely determining the material functions appears to be quite impracticable. Onaran and Findley [46] specialized Lockett's results to simple stress states and presented examples to show the effectiveness of a multiple integral functional relationship for describing the response of polyvinyl chloride. The same approach is used at the end of this chapter. The assumption that the elongation of a simple tension specimen at time t depends on all the previous values of the rate of loading can be expressed mathemati- cally by: t E(t) =6? 3531431] , (4.35) T=-w where E(t) = unit strain of specimen at time t, 0(1) = time-dependent uniaxial stress, F = a functional. Should the functional F be linear and continuous, this response can be represented by Equation (4.12). 101 It has been shown (see [69] for references) that if F is continuous and nonlinear it may be represented to any desired degree of accuracy by: t dO(Tl) E(t) = Dl(t-Tl) ? dTl l t t dG(Tl) do(r2) + D2(t-Tl,t—T2) T7175?— d’Eld’L'Z + o o o t t dO(Tl) + o o o Dn(t-Tl, o o o ,t"Tn) W- o o o do(1n) —a-,-E—— dTl o o o dTn. (4.36) n Spencer and Rivlin [61] showed that for an initially isotropic material the number of such integrals required does not exceed five. For most applications thus far attempted two or three have proven sufficient for accept— able accuracy. If the material has no load history prior to time t = O, the lower limit of the integrals can be replaced by zero. Furthermore, if the stress-—say oi--is suddenly applied at time zero and subsequently maintained constant (creep test), then Equation (4.36) becomes: 102 2 5(t,oi) _ Dl(t)oi + D2(t,t)oi + . . . + Dn(t, . . . ,t)o§. (4.37) The first term is linear in stress; subsequent terms might be regarded as corrections to the linear one to account for nonlinear behavior. A number of investigators, including Onaran and Findley [46], have found three terms of Equation (4.37) sufficient to adequately describe the behavior of several polymers for a reasonably wide range of stress and time. Then: _ 2 3 8(t,Ui) — Dl(t)o’i + D2(t,t)0'i + D3(t,t,t)di. (4.38) Hence loading programs having three independent values of Oi provide three equations for determining the functions Dl(t), D2(t,t), and D3(t,t,t). The function Dl(t) is determined completely, while the other two are determined only where their arguments take on equal values. Though Dl(t) is determined completely, it should be noted [35] that it depends on each of the constants Oi and there- fore will not be identical with the function obtained in the linear theory from a single value of oi. If experimental values of E(tl’oi) for the three chosen values of Oi are used in Equation (4.38), then 103 values for D1, D2, and D3 are obtained only for t1, the time represented by the data. When such values are computed at enough different times, the form of the func- tions Dl(t), D2(t,t), and D3(t,t,t) can be determined. Alternatively, if analytical curves can be fitted to the data for the three chosen values of oi, then analytical expressions can be used in Equation (4.38) to yield the functions Dl(t), etc. directly. Though this approach is somewhat empirical, it extends the purely empirical approach by providing the functions D1(t), 02(t,t), and D3(t,t,t), which are the kernels of the more general "superposition" integrals in Equation (4.36). The problem in this latter approach is to find the analytical expression which satisfactorily fits the data for the three chosen levels of stress. The accuracy of such a fit depends upon the range of stress and time employed, as well as upon the form of the expression and upon the behavior of the material itself-—e.g., upon the "smoothness" of the transition from the linear to the nonlinear region. Onaran and Findley [46] have found an expression of the form 5(t) = so + gth, (4.39) where 104 8 O a curve-fit constant, g,h = constants (to be explained), to yield satisfactory results for several polymers. After rearranging Equation (4.39) and taking the logarithm of both sides, it becomes: log (8 - 80) = log g + h log t. (4.40) Hence on a full logarithmic plot of (8 - 80) vs. t, h is the slope of the line at time t and g is an intercept. The constants 80 may be chosen to make the curve a straight line of best fit, thereby making h a constant. When these constants are determined for three independent stress levels oi, Equation (4.39) can be used with Equation (4.38) to determine the functions (kernels) Dl(t), D2(t,t), and D3(t,t,t). Nonlinear Photoviscoelasticity Virtually all work on nonlinear behavior reported in the literature deals with the mechanical properties of the materials tested. Similarity of the strain—stress and birefringence—stress plots of Chapter III seems to imply that any analytical description of the mechanical behavior would parallel that of the optical behavior——i.e., bire— fringence is very closely related to strain throughout the linear and into the nonlinear range. Observing this and considering birefringence data of this study more accurate than the strain data, it was decided to apply the scheme 105 outlined above to determine similar kernels for an integral-type representation of the photoviscoelastic properties of the material tested. Since axial tension tests were used in this study, questions regarding coincidence of optical and mechanical axes do not enter. In all of the equations of the previous section n(t) might simply be substituted for E(t) where: n(t) = nl(t) - n2(t), (4.41) so that n(t) is the difference of secondary principal indices of refraction--i.e., the relative birefringence or simply the fringe value. Equation (4.38) then implies: n(t o ) = D (t)o + D (t t)o2 + D (t t t)o3 (4 42) ’ i 1 i 2 ’ i 3 ’ ’ i' ° Equation (4.39) suggests: n(t) = no + gth; (4.43) and Equation (4.40) leads to: log (n - no) = log g + h log t. (4.44) The largest stress plotted in the optical test of CR-39 was 2,858 psi. The stress levels 600 psi, 1,600 psi, and 2,600 psi were chosen for the three independent values of stress for use in Equation (4.42). The time range 106 available was one minute to 240 hours (14,400 min). The range one minute to 104 minutes was chosen for representa- tion because of its convenience on the logarithmic time scale. The method of determining nO in Equation (4.44) to cause a "best fit" straight line is not apparent in the literature surveyed. The following scheme was therefore devised. Using data from the curves of birefringence vs. stress at t = constant (Chapter III), a plot of n vs. log was made1 for the three stress levels mentioned above. From this plot values of n were read at one minute, 102 minutes, and 104 minutes at each of the three stress levels. With these three values of n, at a given stress, nO was then computed to force a straight line on the log (n - no) vs. log t plot through the three points. Figure 4.2 shows this plot. Then 9 was evaluated as the ordinate at one minute and h as the slope of the line for each stress2 in accordance with Equation (4.44). This yielded the desired expressions for the left side of Equation (4.42) which may then be written in matrix form, using the values found, as: 1The plot of n vs. t at o = constant in Chapter III does not yield sufficient accuracy at 102 minutes. 2In all publications reviewed the slopes were found to be the same for all values of stress. The variation of h in the present case is evident in the exponents on t in Equation (4.45). 107 .mmlmu mom mam>ma mmmupm mmucu mom wEHp .m> A0: I CV lanes mafia ooo.oH ooo.H ‘ OOH OH MO #OHmllm.v __A__—— _ _—q_—dj— _ _____— CH mm.o EC ofifl ___p_ . E(P _ ...—... _ _p____ q . u u K'U mmlmU _____ _ 108 '— —H r— "r—‘H 2.10 + .28t'1291 600 6002 6003 01 5.86 + .48t'1895 = 1,600 1,6002 1,6003 02 9.77 + .76t’2386 2,600 2,6002 2,5003 D3 . d "' d _"‘(4.45) When the first matrix on the right is inverted, the desired expressions for the kernels are found. The kernels are shown as functions of time in Figure 4.3. It should be noted that D2(t,t) is negative throughout. Greater insight into the contribution of each kernel to the total birefringence is afforded by Figure 4.4, where the three terms on the right side of Equation (4.42) are shown separately, along with their sum and the experimen- tally determined curve of birefringence vs. time at 2,600 psi. The nonlinearity is seen to be quite pronounced; and the deviation between the linear viscoelastic contribution, given by the Dl(t) kernel, and the actual behavior is seen to increase with increasing time. The calculated curve with all three kernels gives excellent agreement with the experimental results. Recall, however, that the stress of 2,600 psi was used in determining the kernels. The accuracy of the calculated values of bire— fringence for a wide range of stress can be judged from Figure 4.5, where the curve of n vs. t (o = const) from Chapter III is shown, along with calculated values of n. Agreement is seen to be quite good (throughout the time 109 .mmnmo mom Au.h.pcmo 6cm Ap.pvmo .Apcflo mamcumxuum.v .mflm \ C ____ H CH mm.o EC Gav _ _ _ _ . Hmpuo N NCH as 60H . Ab.bvmo - ml (5!? OOO.«.O..H. P p . . GOO”H. . p . . . . OOH... . . r . p OH.. . . —._fi.__~ am. _ q 1 . fiATq __ 4 (3...; mmlmu _r_.__ .... _ 110 .mmlmu mom .Hmm oom.m n o .mmsHm> pmpmHSUHoU paw Hmpcmsflummxm .wEHp .m> c umouo wOCflum UHmeounuomHliv.v .mam ml 1%. m lasso mane / ooo.oa ooo.H OOH r o. inx++ i i i i ”iii“ 1 x i x ink; ”4 v x( 4 ..ii ++ . i o -m m b D m -4 To -w baa (OH -NH Hma oom.m Iva ompmHSUHmU -mH HopcmEHquxm n a a ___._ _ . _ _ __.__ r . . _ .7... __ _ . ____. __ t ON u JepJo GBUIJJ : Imu How 0. . . . .monao n o .mmSHm> ompmHSUHmo ocm Hmpcmfiflummxm .mEHv .m> c Hmpuo mmcwum UH#MEOH£UOmHIIm.v .mflm 3.30.: 2:; 0w. ON_ Om — q q _ u — q u _ A _ .....mmduu .Ecozuux .mmnmo 6223.60 . . . BEoEBaxm I .3 006 ..m .2. coo _. b .2. ooo. .b .308. .b 111 ’ (’ - ’ ‘ .ma co! . b .2. cow. . b ...wooom . m .3 comm... . b 7.. OO¢N .- b .naoom_ub ( \\\\\\ .2. coon .b l12 range) for the stress range of 600 psi to 2,600 psi but poor for the stress of 2,858 psi. This is not surprising, since the stresses of 600, 1,600, and 2,600 psi were used to determine the kernels. Furthermore, the nonlinearity becomes very pronounced at stresses above 2,600 psi. This may be observed either in Figure 4.5 or in Figure 3.2 of Chapter III. Closer agreement between experimental and calculated results for the higher stress level of 2,858 psi, of course, is possible by using this value of stress as one of the three for calculating the kernels. However, this would probably cause greater discrepancies at the lower levels of stress. Modifying Equation (4.42) to yield closer agreement with experimental results is also possible. Onaran and Findley suggested the following form: 2 3 n(t,oi) = Dl(t)oi + D2(t,t)(o§) + D3(t,t,t)(oi) , (4.46) where . = — Oi Oi all when Oi2> Ott’ t = Ci 0 when Oi'< GEL. As previously pointed out, the kernel Dl(t) represents the linear contribution to the response, and the other two represent the deviations from linearity. Since this devi- ation is more pronounced at higher levels of stress, the above modification can be expected to improve considerably 113 the representation at these higher levels without loss of accuracy at lower levels, where behavior is essentially linear. Since the linear limit stress 02$ changes signifi- cantly with time, some question arises concerning the proper value to use in Equation (4.46). It seems that the lowest value, corresponding to the longest time to be considered in the representation, should be used. This modified theory was not used for the CR-39 reported here, because the former representation seemed adequate for the stress range of 600 psi to 2,600 psi. For the time range of 240 hours, 2,600 psi is dangerously near the fracture strength of the material. Although it carried a stress of 3,085 psi for the full 240 hours in the optical test, it fractured in less than 167 hours at a stress of 2,858 in the mechanical test. The latter specimen was probably damaged while applying the moire grid, though such damage was not apparent. In earlier optical tests the material fractured in less than 47 hours at 3,700 psi and in less than 2 hours at 4,170 psi. Limitations of Results Presented All results given above are based on constant uniaxial stress sustained for 240 hours. Hence no informa- tion on the behavior of the material in recovery or upon unloading is given, even though such data was collected. 114 Furthermore, this study makes no attempt to determine whether the stress—optic law for more general stress states differs materially from the one for uniaxial stress. On the basis of some published results, significant differ- ences are not expected even in the nonlinear range. For example, in cellulose nitrate Frocht and Thomson [23] found that for equal values of principal stress difference essentially the same birefringence was obtained under a biaxial state of stress as under a uniaxial state, although strains of nearly twenty percent were encountered. As pointed out previously, a thorough investigation of behavior under combined stress states extending into the nonlinear region requires an exhaustive testing program. CHAPTER V VARIATION OF OPTICAL PROPERTIES WITH WAVE LENGTH Investigatioanerformed All results presented in previous chapters involving the relative retardation or birefringent properties of CR—39 were based upon one wave length of radiation-~namely, 410 nm. The photoelastic coefficients, such as CO and f0, in the linear range for this one wave length were deter- mined as functions of time. The extent of the linear ) ) —-was likewise deter— range—-i.e., or 1.0 n(OM, 5.0 mined as a function of time for one wave length. n(ozz Most optical material properties depend to some degree upon wave length of radiation. An extreme example is the high transmission of some wave lengths and complete blocking out of others, a phenomenon useful in the produc- tion of filters and in x-ray techniques, to name only two familiar cases. The birefringent properties are likewise known to depend upon wave length. This dependence is known as dispersion of birefringence. Since a number of wave lengths, ranging from 410 nm to 800 nm, were employed in this investigation, data on this phenomenon was obtained. Further discussion and some data on dispersion of bire- fringence are given in the following section. 115 116 In view of the dependence of birefringence on wave length, it seems logical to question whether the linear limit stress at a given deviation from linearity-~say n(dzz)l.0-'might also depend upon wave length. This was, in fact, the primary purpose of using several wave lengths. Although the results of this investigation are not conclu- sive in answering the question, some useful results were obtained and are presented in the last section of this chapter. Dispersion of Birefringence The time-dependence of the photoelastic coefficients in Equations (2.7)-(2.9) was previously discussed. Disper- sion of birefringence implies that the coefficient CO is not independent of wave length, nor is the coefficient fO directly proportional to wave length. It is equivalent to say that the fringe order n is not inversely proportional to wave length. Dispersion of birefringence may be regarded as a source of error, a nuisance, or an asset, depending upon its recognition and utilization. If several wave lengths are employed in a photoelastic investigation and the dis— persion of birefringence ignored, significant errors may result even though a relatively narrow range of wave lengths is employed. Such errors, of course, can be avoided by careful calibration tests employing the same wave lengths used in the investigation, resulting then in 117 some inconvenience. For at least one material, dispersion of birefringence was successfully used to measure the degree of plastic deformation in photoplasticity investiga— tions [42]. In this case dispersion might well be considered an asset. Several types of measurements and descriptions of dispersion of birefringence have been employed. Coker and Filon [6] reported the results of their many investigations on glass and referenced the work of several other investi- gators. Though their methods of measurement varied consid- erably, results obtained were sufficiently consistent that a description of dispersion of birefringence was possible. For many glasses they found that the variation with wave length of the stress-optic coefficient CO was adequately described by: C = , (5.1) where CO and 10 are constants for a given glass. The authors pointed out that "the formula is not to be regarded as of universal application" and that even some glasses must be considered outside the range of applicability of the formula. Monch [41] expressed dispersion of birefringence as a percentage by the equation: 118 (n1)Na - (n1) Hg D = . 100, (5.2) (n1)Na where (n1)Na and (n1)Hg are the products of fringe order and wave length for yellow sodium (1 = 589.3 nm) and blue mercury (l = 435.8 nm) radiation, respectively. In later work Monch and Loreck [42] substituted red light (A = 655 nm) for the yellow sodium above, resulting in greater values of D and therefore greater accuracy. Although the above description with either value served the desired purpose-—namely, representing the degree of plastic defor— mation in celluloid——the arbitrariness of the choice of wave lengths employed should be noted. No insight is given into the fundamental nature of dispersion of birefringence at any wave lengths other than those used in the descrip— tion, covering a rather narrow range. The shortcomings of previous descriptions as general measures of dispersion of birefringence were discussed by Pindera and Cloud [49]. A new description, called "Normalized Dispersion of Birefringence," was developed which removed most of the shortcomings of those previously presented. Perhaps most significant, it allows dispersion of birefringence to be determined as a continuous function of wave length. The procedure outlined below follows closely that given by the authors cited. Some 119 modifications, reflecting mostly personal preference, are introduced in the following. In the creep test performed on CR-39 a sequence of pictures at various wave lengths was taken, beginning at 100 hours after the constant load was applied. Since the entire sequence was obtained in about six minutes, creep effects were negligible. From these pictures data was obtained for the plot of fringe order vs. length along the tapered specimen. Since the stress was known as a contin- uous function of the length, fringe order could then be plotted as a function of stress at each of the wave lengths. This plot is shown in Figure 5.1 for eight wave lengths.1 Then at o = 1,200 psi (within the linear range of behavior) the fringe order n was read for each wave length. These values and the "relative retardation" R (R = ml) were tabulated. In the linear range these values are, of course, linear functions of stress. The ratios _ R(l) r -rn'57’ (5'3) where 10 = 546 nm is an arbitrarily chosen "reference" wave length, were then computed and tabulated. 1Pictures at 500 nm and 700 nm exhibited irregular intensity fringes and therefore were not used. 120 .mmlmu mom ax. . . . .mxnax n K .mmwnpm .m> c umpuo mosanm UHDDEOMSUOmHIIH.m .mfim :2: b OO .vd ooo.~ Dow; 00 N._ 00m 00¢ - fi 1 a q .536 no ”.09 1 .mm . mo 4\\\,///, x / , / .5. com ... / , .52: L / ////| gE.ommuA . / / .5: 00m u A .6: Fun "A / ...... mg "4 .E: wnv "A .6: 0;. «A d d d d d obugxg Q 9 u ‘10910 t. m. 121 The ratios of Equation (5.3) were plotted as a function of wave length. Pindera and Cloud called this a "Normalized Retardation Curve." The rate of change of normalized retardation with respect to wave length was then defined as the "Normalized Dispersion of Birefringence" and denoted by D Thus x. _ dr DX _ 8")?" (5.4) Both r and D as defined by Equation (5.3) and 1’ Equation (5.4), respectively, are independent of stress in the linear range but may well depend upon stress in the nonlinear range. This might be expected on the basis of the previously mentioned results for celluloid. In fact, the dependence of r on stress could seemingly be an indication of nonlinearity. To investigate this possibility for CR—39, normalized retardation curves at one level of stress in the linear range (1,200 psi) and several levels in the non— linear range were plotted. Figure 5.2 shows a reduction of such a plot in which only two levels of stress in the non— linear range are included. Some variation with stress is evident. The curve for the stress in the linear range is Similar in many respects to the one given by Cloud [4] for CR—39. Cloud found no systematic variations with uniaxial Stress: variations were random. He analyzed his results 122 .mmlmU MOM mm>HSU COHpMUMMQOH UONHHMEMOZIIN.m confirm. $0.0 123 statistically and showed a range of r values at each wave length employed. Results of this study are subject to significant error because of the sensitivity of r to wave length and the lack of precise determination of the wave lengths used. Manufacturer's specifications of filters employed were checked using a hand spectroscope (Zeiss) and the values were accepted. The slope of the normalized retardation curve in the linear range of stress was evaluated at several points. From these slopes the normalized dispersion of birefring— ence curve, Figure 5.3, was plotted. Again some similari- ties to the corresponding curve obtained by Cloud for CR-39 are evident, as well as some striking differences. The dispersion found by Cloud was negative throughout the range of wave lengths employed at 66 hours after the load was applied. Although the increased time of 100 hours used in this study could account in part for the differences, it could not be solely responsible for them. Further inves- tigations seem necessary to reconcile the differences. In an attempt to correlate dispersion of birefring- ence phenomena to the degree of nonlinearity of birefring- ence, a plot of normalized retardation vs. stress was constructed for each wave length. Readings were taken at several values of stress in the nonlinear range. However, since no conclusions could be drawn from the plot, it is 124 .mmlmu mom DUCDWCHMMDHHQ mo coamumdmflo pmNHHMEMOZIIm.m cowfib M mmlmu .3 O O H II 00m N+ 125 not shown. Figure 5.2 provides a glimpse of the difficulty encountered. The curves for 2,000 psi and for 2,800 psi are hardly distinguishable from each other, despite wide differences in the degree of nonlinearity. It must be concluded that the normalized retardation and hence the normalized dispersion of birefringence are not effective measures of the nonlinearity of CR-39, even though they cannot be assumed independent of stress. The magnitude of the error committed in assuming that photoelastic coefficients are independent of wave length (in the linear range) can be estimated from Figure 5.2. The possible error in the visible spectrum (390 nm to 770 nm) for CR-39 appears to be about seven percent. Variation of Linear Limit Stress with Wave Length Figure 5.1 readily yields data on the dependence of the linear limit stress on wave lengths. This is presented ) ) and in Figure 5.4, where 1.0 n(ozg are given for n(ou 5.0 each wave length. Considerable variation of the linear limit stress values may be observed. Values of ) are within mm” 1.0 about ten percent of a "mean" value of 1,355 psi; values of n(dzz)5.0 are Within seven percent of 1,720 p51. These variations seem quite random, however, perhaps indicating only the magnitude of experimental error rather than any 126 .mmtmu Mom mummmuwm uHEHH HDQCHH c mcflsogm . A. .m> s umouo mmcflum UHmeOHnuolelw.m 00 ¢.N 000.N 000.. W . 4 . . A . . . . . .536 "u “.09 1 .3 (mo / / / .5: one ...A / // .8: com n K ////.E: @VD "K o o o nNKnHJA :2: 6 com; q q u d 4 Seen» .6: ppm "A .E: wnv u A .52.. O_¢ "K n A .mwmhum .mflm fi , . ...... com... A com ‘0 (w u ‘ JOpJO Ghana 9 ¢. w. m. 127 dependence on wave length. Such dependence cannot be detected in the results of this study. Causes and consequences of the irregularities of linear limit stress values are further discussed in the next chapter. It might here be mentioned that some irregu- larities must be expected. Limiting-value types of material properties (such as proportional limit, endurance limit, or ultimate strength) are not easily determined. Usually a large number of specimens is tested and the data is subjected to a statistical analysis. Precise values of such properties cannot be obtained from a single specimen. CHAPTER VI DISCUSSION, CONCLUSIONS, AND RECOMMENDATIONS FOR FURTHER RESEARCH Achievement of Objective The primary objective of this study——to investigate the effect of time upon the birefringent and mechanical properties of several plastics--was achieved. The iso- thermal, quasi-static behavior under constant, uniaxial stress conditions was examined quite exhaustively--for approximately five decades of logarithmic time, and for a range of stress sufficiently broad to cover not only the linear region but also much of the nonlinear region for those materials exhibiting nonlinear behavior. Data is presented graphically for all materials. Certain characteristic functions defined by viscoelastic and photoviscoelastic theories are determined for CR—39 in both the linear and the nonlinear regions. Such charac— teristic functions can be deduced from data presented in Appendix III for the other materials tested. The linear limit stresses--i.e., the stresses limiting the momentarily linear range of behavior—-were determined as functions of time for the response to stress of both strain and birefringence. The linear limit 128 129 stresses at both one percent and five percent deviation from linearity were determined. For CR—39 these stresses are plotted as functions of time in Chapter III. They are shown on birefringence-stress and strain—stress plots in Appendix III for the other materials. Further discussion of specific properties of each material appears later in this chapter. A number of wave lengths of radiation were employed in the birefringence investigations, providing data for the study of two phenomena: dispersion of birefringence and variation of linear limit stresses with wave length. Such data was analyzed only for CR—39. Experimental Techniques Tapered models were successfully employed to obtain optical and mechanical data essential to this study. Few references can be found to the use of such models by other investigators. Their use in conjunction with the moire method for collecting mechanical strain data is believed to be without precedent. Tapered models offer the advantage over most commonly used calibration models of allowing measurement of the response to stress over a broad range of stress. In principle, the response can be measured as a continuous function of stress within the existing range. This avoids the necessity of testing a large number of models or of using repeated or step loadings (which would be necessary 130 for prismatic models) to obtain data such as that required by this study. Because stresses near the ultimate strength were reached, neither repeated nor step loadings would have been appropriate. The use of tapered models thus made feasible a very extensive investigation into the effects of time on the materials tested. More general use of such models for studies of material behavior is recommended. Although the models and techniques employed yielded results acceptably accurate for this study, some improve- ment in models and techniques is possible. The equipment available for this study would have allowed longer models. Somewhat less than a lO—inch—long gage portion of both the photoelastic and the mechanical specimens was actually useful because of the interference produced by the load in the stress distribution near the ends. The fields of both the polariscope and the optical system employed in observ— ing moire fringes were thirteen inches in diameter. The use of longer models would permit a slightly more accurate determination of stresses and would reduce the significance of errors in determining fringe locations. However, this improvement would be very small in the present case. A small improvement might also result from using higher—density line arrays in the moire analysis. A suffi— Ci‘entnumber of points was available for plotting very SI“Both curves of fringe order vs. length. But higher— dearlsity arrays would yield not only a larger number but 131 also narrower fringes, permitting more accurate fringe location determinations. The same advantage could be gained by using an optical sensing device in conjunction with an x—y recorder, as described in Appendix II. Improvement in technique is needed chiefly in slope determinations of the moire fringe order vs. length curves. Better results might be obtained by use of an analytical (numerical) approach wherein smooth curves are fitted to the experimental data points and the slopes of these curves are evaluated numerically at desired points to yield the desired strains. These strains could then be related to stresses at the points where strains are known and stress- strain curves could be plotted. Birefringence-stress curves could be plotted in much the same way. From these curves linear limit stresses for any chosen deviation from linearity could be evaluated. It seems that the most reliable check of the over-all accuracy of such an ambitious numerical program would be a comparison with results obtained by the proce— dures followed in this study. The proposed numerical approach might conceivably reduce the irregularity of data point locations on some plots, such as those for strain vs. stress or linear limit stress vs. time. The effects of such irregularity were minimized by the graphical tech— niques employed in this study. It was felt that the increase in accuracy, if any, which might have resulted 132 from using a numerical approach rather than a graphical one, could not justify the expenditure of time required by both the computer and the programmer to pursue such an approach. The results of the study of dispersion of bire- fringence in CR-39 are not conclusive, even though they compare quite favorably with some results previously pub— lished by other authors. The main source of error is believed to be in the determination of the wave lengths-—a very crucial determination for such studies. A hand—held spectroscope was used to check the manufacturer's specifi- cations of the filters used. Although no deviations from the Specifications were detected, the sensitivity of the spectroscope readings (about i 2 nm) was not considered adequate for the purpose at hand. Some variation with stress of the normalized retardation curve was detected. The variation was not con- sistent enough, however, to indicate a definite relation- ship between normalized retardation and stress, or between normalized retardation and degree of nonlinearity in the birefringence response to stress. No dependence upon wave length of the linear limit stresses could be detected. There is probably no good reason to expect such dependence, at least within the limited range of wave lengths employed. However, this part of the study yielded valuable information regarding the 133 magnitude of possible error in all previous determinations ) of linear limit stresses. Apparently the n( values 022 1.0 found for CR-39 might be in error by as much as ten percent. But this does not mean that a serious error would result from assuming linearity to extend to a stress level that is ten percent higher than the actual value of n(ozz)1.0° It simply means that the deviation from linearity of the response (birefringence or strain) might be, say, two percent rather than one percent. Hence the errors which may be present in the reported values of the linear limit stresses do not seriously limit the usefulness of the results of this study. No serious errors will result from assuming the behavior of the materials to be linear up to the linear limit stresses reported, even if these values are not pre- cisely determined. Serious errors can result, however, from assuming that a material such as CR—39 exhibits linear behavior up to the ultimate strength. Hopefully this point has been made eminently clear by this study. Material Property Descriptions One fundamental aspect of the forms of material property descriptions and of their interpretation deserves further discussion. Suitable viscoelastic and photovisco- elastic forms of representation were discussed in Chapter IV. In Chapter II it was implied that the characteristic functions describing the mechanical properties might be 134 thought of as time-dependent elastic coefficients and might be used in a modified, time-dependent Hooke's law. The strain response to one-dimensional constant stress was given as: E(t) = m. (6.1) The quantity E(t) was called a relaxation modulus, but the resemblance to a time—dependent Young's modulus was pointed out. Similarly, the commonly used birefringence-stress relations were stated as: n = 42%— (C51 - 02) (6.2) O and Cod It was shown that the coefficients f0 and CO actually vary with time. Again in this case it was implied that "momentary" values of these coefficients, as deter- mined by one—dimensional tests, can be used in the more general stress situations represented by Equations (6.2) and (6.3). Thus a "momentary—photoelastic" condition of the material is implied, although the materials tested are admittedly photoviscoelastic. The accuracy of many previously conducted photo— elastic investigations, using materials such as those 135 tested in this study, indicates that the assumption of momentary "photoelasticity" is valid. This furthermore suggests that the assumption of momentary elasticity is valid in the case of mechanical properties. In this study no serious effort was directed toward establishing a firm theoretical basis for such assumptions, nor was an intensive search made of the literature for possible establishment by other investigators. Some ideas or first impressions are believed to be worth stating and are therefore given below. In the discussion of viscoelasticity (Chapter IV) the relationship between the relaxation modulus E(t) and the extension compliance D(t) was given as: t J/~ E(t—T)D(T)dT = t. (6.4) 0 But for "moderate rates of cree " an a roximate relation PP is: E(t) ; DTET’ (6.5) Then the one-dimensional strain response to stress is given by Equation (6.1). But this equation assumes a modified form of "Hooke's law"; the relaxation modulus E(t) is con- sidered simply to be a time-dependent Young's modulus. Hence the accuracy of the approximation involved in 136 Equation (6.5) is believed to be closely (if not directly) related to the momentary-elastic assumption. Some insight into the problem might be gained by approaching it from another point of view. Differentiating with respect to time the one-dimensional Hooke's law equation Q ll Es, (6.6) we obtain: Es + Es, (6.7) Q II where the dots denote time derivations and E is Young's modulus. Now if the behavior is truly elastic, E must be zero, since E is a constant. Then Equation (6.7) becomes: 0 6: E8. (6.8) For a momentary-elastic material the sequel to Equation (6.8) must be: 6:. D(t)§:. (6.9) But by starting with a time—dependent Hooke's law 0 = E(t)8, (6.10) we obtain by differentiation: 6: E(t)s + D(t)é. (6.11) For Equation (6.11) to be identical to Equation (6.9) we 137 must have: E(t) E 0. (6.12) Thus if the time-dependent Hooke's law assumption is to yield satisfactory results, the condition E(t) 3'0 (6.13) must prevail. This condition is implied by the "moderate rates of creep" condition leading to Equation (6.5). Whether or not the two are more rigorously and fundamen— tally related, both are believed to be applicable to the behavior of the materials tested in this study. Properties of Particular Materials Polyester Resin CR-39.—-The properties of this material were discussed from many viewpoints in Chapters III-V. It was selected for testing and for extensive discussion because it is probably the most widely used photoelastic material in this country. It has many desir- able properties: It is clear and very smooth-surfaced, thus possessing excellent transmission properties and yielding very distinct, regular fringes. Furthermore, it is in an intermediate range of optical sensitivity and exhibits linear behavior throughout a reasonably wide range. It is inexpensive and readily available in sheets of various sizes and thicknesses. It will undoubtedly continue to be widely used. 138 However, this material must be carefully calibrated, not only for the effects of time upon the photoelastic coefficients and the mechanical characteristic functions but also for the extent of the linear range of behavior. This is particularly important because CR-39 is available from a number of suppliers and its properties are known to vary with the manufacturer. The results of this study show that the linear limit stresses vary greatly with time after loading. Further» more, within a few minutes after load is applied the linear limit stresses are much lower than the ultimate strength. This makes an analysis based on the assumption of linearity up to near-failure vulnerable to serious error, since large deviations from linearity are clearly possible. Polyester Resin PS-l.=~Although this material is intended primarily for photoelastic coatings, its proper~ ties make it very useful for models also. The results obtained from the tests on this material are presented graphically in Figures A3.1-A3.5 of Appendix III. The material is seen to exhibit a distinct linear range of behavior, both mechanical and optical. The change of linear limit stress with time is small. The values of n(d ) are lower than the corresponding values RE 1.0 for strain 5(Gflz)l.0° reported for CR—39. This is the same result herein 139 Deviations from linearity are relatively "slow": that is to say, errors would be small in assuming linearity to extend well beyond the linear limit stresses reported. Furthermore, the birefringence-strain plot (Figure A3.5) is very nearly linear up to the highest strains encountered, about two percent. This is a very important property in regard to the material's potential as a photoelastic coating. Figures A3.2 and A3.4 show that time effects are extremely small after the load has been applied for three or four hours——i.e., the creep rate is very low. The material is very nearly clear; some roughness of the surface detracts from its otherwise good transmission characteristics. It is very sensitive optically and is easily machined. It was found to yield (not fracture) at a stress above 9,000 psi. Therefore, it may be a useful material for photoplasticity studies. Its lack of brittle— ness reduces the hazards normally associated with stress concentrations. Over-all, PS-l is considered to be an excellent photoelastic model and coating material. Epoxy Resin PS-2.——The behavior of PS—2 is similar in many respects to that of CR—39. It exhibits a much lower creep rate, however, and a smaller variation with thma of linear limit stresses. Curves showing the data Obteiined for this material appear in Figures A3.6-A3.10. 140 No deviation from linearity in either the optical or mechanical response to stress was detected within the first few minutes after load was applied. This is contrary to earlier findings of the optical behavior as reported in [50]. A different lot number of the material was used in this study than in the earlier investigation. The linear limit stresses at one percent deviation from linearity for optical behavior are slightly lower than the corresponding ones for mechanical behavior. The creep rate of PS-2 is somewhere between those of CR—39 and PS—l. Figure A3.10 shows that the birefringence- strain relation is linear for a wide range of strains (above 1%) and deviates very little from linearity for strains up to nearly two percent. Like PS-l, it is very useful for coatings. Birefringence is seen to decrease with time at a given strain. The material is ductile: it fails in shear at a uniaxial tensile stress above 12,000 psi. It has a deep amber color'which presents some problems concerning fringe definition (low contrast). It has very smooth surfaces and is very sensitive optically (comparable to PS—l). Like PS—l, it is available in sheets of various sizes and thicknesses. Polyester Resin Palatal P6-K.--Figures A3.11-A3.15 show the optical and mechanical properties of P6-K. No deviation from linearity was detected in the optical 141 behavior within the ranges of stress and time employed. Some nonlinearity was observed in the mechanical behavior, but the deviations from linearity are seen (Figure A3.13) to be small and to occur only after constant stress was maintained for several hours. The Specimens used for the two investigations (optical and mechanical) were from different plates which were cured separately. Although the plates were taken from the same "lot" and approximately the same curing cycle was used, the differences in the behavior (mechanical and Optical) might be due to variations in the handling of the specimens. The creep rate is very low. Birefringence is seen (Figure A3.15) to decrease rapidly with time at a given strain. The material was donated by the manufacturer in liquid form. It was mixed in the following proportions by weight: 100 parts resin, 1.50 parts catalyst, and 0.05 parts activator. It was cast between glass plates in sheets of nominal l/4—inch and 3/8—inch thicknesses. A curing temperature of 90°C was maintained for 40 hours. The material was tested after it was stored, at room temperature, for over a year. Though relatively insensitive optically, the broad range of linearity in the birefringence response to stress makes it a very useful photoelastic model material. 142 Fracture occurred in the photoelastic specimen after 200 hours at a stress of 5,950 psi. The test for mechani- cal properties was terminated after 100 hours because of an air conditioner failure. Palatal P6-K exhibits good optical transmission characteristics with only a slight amber tint. It is quite brittle and is susceptible to chipping during the machining process. Recommendations for Further Research Several suggestions for continuing or extending the work of this study seem in order. These are given below: 1. Constant strain, relaxation-type tests, both optical and mechanical, should be conducted to check some results deduced from the creep tests of this study. Such tests on CR—39 would be particularly enlighten— ing in View of the possible inflection in the birefringence—strain plot (Figure 3.8). A direct check could be made on the moduli, compliances, and photoelastic coefficients reported in this study. 2. Only quasi-static loadings were employed in this investigation. Yet accurate calibration tests are equally or even more important to dynamic studies using models. Some experimental studies of wave propagation, for example, have been reported with wholly inadequate calibrations, or even a complete 4. 5. 143 absence of them. Thus, the short-time response to periodic and/or impact loadings should be accurately determined. Constant strain rate loadings might also be employed. Such studies would constitute a valuable extension to the results here reported, if the same materials were used. Photoviscoelasticity promises to become a powerful tool in viscoelastic stress analysis. An intensive search for suitable model materials must be con— ducted if the method is to achieve prominence. These materials must, of course, be carefully exam- ined (calibrated). Photoviscoelastic theories and techniques must be scrutinized, tested, and further developed. The momentary-elastic assumption, discussed at various points throughout this writing, seems justi— fied for high—modulus materials exhibiting a low creep rate. But the question remains open of when (for what creep rate, say) the momentary—elastic theory must be abandoned in favor of a more general viscoelastic theory. Theoretical and/or experi— mental studies to help answer this question would be very valuable. Linear limit stresses and certain other characteris- tic functions describing material behavior were determined in this study for the response to 144 one-dimensional stress. The ability of these quantities or functions to predict behavior under more general stress states should be checked. 6. Little experimental evidence has been provided for the study of nonlinear viscoelasticity under com- bined stress states. A number of theories and assumptions, therefore, lack experimental verifica- tion. This is particularly true for nonlinearities associated with small strains. Nonlinearities of this type are obviously exhibited by some of the materials (e.g., CR—39) tested in this study. Further investigations, employing combined stress states extending into the nonlinear range, might provide valuable data for further theoretical and experimental studies. The moire method of strain (displacement) analysis has been the subject of very intensive study during the past few years. Further development is believed needed, particularly in the techniques of grid production and application. The possible use of diffraction gratings, especially, should be further investigated and developed. 10. 11. LIST OF REFERENCES Alfrey, T., Mechanical Behavior of High Polymers, vol. VI, Interscience Publishers, Inc., New York, 1948. Bell, J. F., "Diffraction Grating Strain Gage," Proc. Soc. Exptl. Stress Anal., 11, 2 (1960), 51-64. Clark, A. B. J., "Static and Dynamic Calibration of a Photoelastic Model Material, CR-39," Proc. Soc. Exptl. Stress Anal., 14, l (1956), 195—204. Cloud, G. L., "Infrared Photoelasticity: Principles, Methods, and Applications," Ph.D. dissertation, College of Engineering, Michigan State University, 1966. Clutterbuck, M., "The Dependence of Stress Distribution on Elastic Constants," Brit. Jour. Applied Phys., 9, 8 (1958), 323-29. Coker, E. G., and L. N. G. Filon, A Treatise on Photo- elasticity, The University Press, Cambridge, Gr. Brit., I957. Coolidge, D. J., Jr., "An Investigation of the Mechani- cal and Stress-optical Properties of Columbia Resin, CR—39,".Proc. Soc. Exptl. Stress Anal., 6, l (1948), 74-820 - Crisp, J. D. C., "The Measurement of Plane Strains by a Photoscreen Method," Proc. Soc. Exptl. Stress Anal., ‘15, l (1957), 65. Dally, J. W., and W. F. Riley, "Experimental Stress Analysis with the Moire Fringe Method," Product Eng., .36, part 2 (1965), 41-48. Daniel, I. M., "Quasi—static Properties of a Photo- viscoelastic Material," Exptl. Mech., 5, 3 (1965), 83—890 '- Daniel, I. M., "Static and Dynamic Stress Analysis in Viscoelastic Materials," Ph.D. dissertation, Illinois Institute of Technology, 1964. 145 12. 13. 14. 15. 16. 17. 18. 19. 20. 21. 22. 23. 146 Diruy, M., "L'Analyse des Contraintes par la Methods des Reseaux Optiques," Docaero, no. 55 (March, 1955), pp. 1‘18. Dolan, T. J., and W. M. Murray, "Photoelasticity," Handbook of Experimental Stress Analysis, ed. M. HeEéInyi, John Wiley and Sons, New York, 1950. Douglas, R. A., C. Akkoc, and C. E. Pugh, "Strain- field Investigation with Plane Diffraction Gratings," Exptl. Mech., 5, 7 (1965), 233-38. Durelli, A. J., and V. J. Parks, "Moire Fringes as Parametric Curves," presented at 1966 S.E.S.A. spring meeting, Detroit, Michigan, May 4-6, 1966. Eirich, F. R., Rheology, Theory and Applications, vols. I, II, and III, Academic Press, New York, 1960. Ferry, J. D., Viscoelastic Properties of Polymers, John Wiley and Sons, New York, 1961. Frocht, M. M., Epotoelasticity, vols. I and II, John Wiley and Sons, Inc., New York, 1941. Frocht, M. M., and Y. F. Cheng, "An Experimental Study of the Laws of Double Refraction in the Plastic State in Cellulose Nitrate-~Foundations for Three—dimensional Photoplasticity," Photoelasticity, ed. M. M. Frocht, Pergamon Press, New York, 1963, pp. 195—216. Frocht, M. M., and Y. F. Cheng, "On the Meaning of Isoclinic Parameters in the Plastic State in Cellulose Nitrate," Jour. Applied Mech., 22 (1962), 1-6. Frocht, M. M., and R. Guernsey, Jr., "Further Work on the General Three—dimensional Photoelastic Problem," Jour. Applied Mech., 22, 2 (1955), 183-89. Frocht, M. M., and R. A. Thomson, "Further Work on Plane ElastOplastic Stress Distributions," Photo- elasticity, ed. M. M. Frocht, Pergamon Press, New Frocht, M. M., and R. A. Thomson, "Studies in Photo- plasticity," Proc. Third U.S. Natl. Cong. Applied Mech. (1958), 533-40. 24. 25. 26. 27. 28. 29. 30. 31. 32. 33. 34. 35. 147 Green, A. E., and R. S. Rivlin, "The Mechanics of Nonlinear Materials with Memory," Archive for Rational Mech. and Anal., Part I, l (1957), 1—21; Part II, with A. J. M. Spencer, 3 (1959T, 82-90; Part III, 4 (1959- 60), 387-404. '" " Gross, B., Mathematical Structure of the Theories of .Xiscoelasticity, Hermann, Paris, 1953. FSee also: Jour. Applied Phys., 18, 212 (1947); 19, 257 (1948); and 22, 1035 (1951). “' “' Guild, J., The Interference Systems of Crossed Dif- fraction Gratings, CIarendon Press, Oxford, 1956. Gurtin, M. E., and E. Sternberg, "On the Linear Theory of Viscoelasticity," Archive for Rational Mech. and Anal., 11, 4 (1961), 291—356. Jessop, H. T., and F. C. Harris, Photoelasticity Principles and Methods, Dover PubliCations, New York, I949. ' "Kodak Wratten Filters," Scientific and Technical Data Book B—3, Eastman Kodak Co., Rochester, N.Y., 1965. Kolsky, H., "Experimental Studies of the Mechanical Behavior of Linear Viscoelastic Solids," Technical Report No. AM—18, Dept. of Defense, ARPA, Materials Research Program, April, 1965. Leaderman, H., Elastic and CreepProperties of Fila- mentous Materials, Textile FoundatiEn, Washington, 1943. Leaderman, H., "Viscoelastic Phenomena in Amorphous High Polymeric Systems," Rheology, vol. II, ed. F. R. Eirich, Academic Press, New York, 1958. Lee, E. H., "Viscoelastic Stress Analysis," Proc. First Symposium on Naval Structural Mech., Pergamon Press, New York, 1960, pp. 456-82. Lee, E. H., and T. G. Rogers, "Solution of Visco— elastic Stress Analysis Problems Using Measured Creep or Relaxation Functions," Jour. Applied Mech., 30 (1963), 127-33. '__ Lockett, F. J., "Creep and Stress—relaxation Experi- ments for Non—linear Materials," Int. Jour. Eng. Sci., '3 (1965), 59-75. 36. 37. 38. 39. 40. 41. 42. 43. 44. 45. 46. 47. 148 Low, I. A. B., "The Moire Method Using Diffraction Gratings," presented at 1966 S.E.S.A. spring meeting, Detroit, Michigan, May 4-6, 1966. Low, I. A. B., and J. W. Bray, "Strain Analysis Using Moire Fringes," Engineering, 213 (1962), 566. Marin, J., and J. E. Griffith, "Creep Relaxation of Plexiglas IIA for Simple Stresses," Jour. Eng. Mech. Div., Proc. A.S.C.E., 82, EM—3 (1956), Paper 1029, 1-20. ‘- Mindlin, R. D., "A Mathematical Theory of Photovisco- elasticity," Jour. Applied Phys., 22 (1949), 206-216. Mindlin, R. D., "Analysis of Doubly Refracting Mate- rials with Circularly and Elliptically Polarized Light," Jour. Optical Soc. Am., 31 (1937), 288. Ménch, E., "Die Dispersion der Doppelbrechung bei Zelluloid als Plastizitatsmas in der Spannungsoptik," Zeitschrift fur angewandte Physik, g, 8 (1954), 371-75. M6nch, E., and R. Loreck, "A Study of the Accuracy and Limits of Application of Plane Photoplastic Experi- ments," Photoelasticit , ed. M. M. Frocht, Pergamon Press, New York, 19* , pp. 169-84. Morse, 5., A. J. Durelli, and C. Sciammarella, "Geometry of Moire Fringes in Strain Analysis," Jour. Eng. Mech. Div., Proc. A.S.C.E., EM—4 (1960), pp. 105- 126. Nielson, L. E., Mechanical Properties of Pol mers, Reinhold Publishing Corp., Chapman and Hall, L ., London, 1962. O'Haven, C. P., and J. F. Harding, "Studies of Plastic Flow Problems by Photo-grid Methods," Proc. Soc. Exptl. Stress Anal., 2, 2 (1944), 59-70. Onaran, K., and W. N. Findley, "Combined Stress-creep Experiments on a Nonlinear Viscoelastic Material to Determine the Kernel Functions for a Multiple Integral Representation of Creep," Trans. Soc. Rheology, 9, 2 (1965) , 299-327. " Pindera, J. T., "Einige rheolgische Probleme bei spannungsoptischen Untersuchungen," Internationales spannunQSOptischen Symposium, Berlin, April, 1961, Akademie-Verlag, Berlin, 1962. 48. 49. 50. 51. 52. 53. 54. 55. 56. 57. 58. 149 Pindera, J. T., "Remarks on Properties of Photovisco- elastic Model Materials, Exptl. Mech., 6, 7 (1966), Pindera, J. T., and G. L. Cloud, "On the Description of Dispersion of Birefringence of Photoelastic Mate- rials," Exptl. Mech., 6, 9 (1966), 470-80. Pindera, J. T., and E. W. Kiesling, "On the Linear Range of Behavior of Photoelastic and Model Mate- rials," Proc. Third Int. Cong. EXptl. Stress Anal., West Berlin, March, 1966, VDI-Berichte No. 102, 1966, pp. 89-940 Post, D., "The Moire Grid-analyzer Method of Strain Analysis," Exptl. Mech., 5, 11 (1965), 368—77. Ross, B. E., C. A. Sciammarella, and D. Sturgeon, "Basic Optical Law in the Interpretation of Moire Patterns Applied to the Analysis of Strains--Part 2," Exptl. Mech., 5, 6 (1965), 161—66. Sciammarella, C. A., "Basic Optical Law in the Inter— pretation of Moire Patterns Applied to the Analysis of Strains-~Part 1," Exptl. Mech., 5, 5 (1965), 154-60. Sciammarella, C. A., "Techniques of Fringe Interpola- tion in Moire Patterns," presented at Second Int. Cong. Exptl. Mech., Washington, D.C., September 28- October 1, 1965. Sciammarella, C. A., and F. Chiang, "The Moire Method Applied to Three-dimensional Elastic Problems," Proc. S.E.S.A., 4, 11 (1964), 313-19. Sciammarella, C. A., and A. J. Durelli, "Moire Fringes as a Means of Analyzing Strains," Jour. Eng. Mech. Div. , PI‘OC. A.S.C.Eo , _8_Z, EM‘l (1961) , 582-601. Sciammarella, C. A., and N. Lurowist, "Interpolation of Moire Fringe Orders by a Purely Optical Technique, Moire Fringe Multiplication," presented at Fourth Int. Cong. Applied Mech., Minneapolis, Minnesota, June 14- 17, 1966. Sciammarella, C. A., and D. Sturgeon, "Substantial Improvements in the Processing of Moire Data by Optical and Digital Filtering," Proc. Third Int. Cong. EXptl. Stress Anal., West Berlin, March, 1966, VDI- Berichte No. 102, 1966. 59. 60. 61. 62. 63. 64. 65. 66. 67. 68. 69. 150 Shepherd, R., and L. M. Wensley, "The Moire-fringe Method of Displacement Measurement Applied to Indirect Structural—model Analysis, Exptl. Mech., 5, 6 (1965), 167-76. Shimada, H., and M. Chiba, "Photoelastic Investigation of Large Deflections by Low-modulus Materials," Exptl. Spencer, A. J. M., and R. S. Rivlin, "Further Results in the Theory of Matrix Polynomials," Archive for Rational Mech. and Anal.,_4 (1959—60), 214—30. Staverman, A. J., and F. Schwarzl, "Linear Deformation Behavior of High Polymers," Die Physik der Hochpoly- meren, vol. IV, ed. H. A. Stuart, Springer-VerIag, Berlin, 1956, chaps. l and 2. Stuart, H. A., Die Physik der Hochpolymeren, vol. IV, Springer-Verlag, BerIin, I956. Theocaris, P. 8., "A Review of the Rheo-optical Pro— perties of Linear High Polymers," Exptl. Mech., 5, 4 (1965), 105-114. Theocaris, P. 3., "Creep and Relaxation Contraction Ratio of Linear Viscoelastic Materials," Jour. Mech. Phys. Solids, 12 (1964), 125-38. Theocaris, P. S., "Moire Fringes: A Powerful Measur- ing Device," Applied Mech. Revs., 15, 5 (1962), 333-39. '—- Theocaris, P. S., and H. H. Kuo, "The Moire Method of Zonal and Line Gratings," Exptl. Mech., 5, 8 (1965), Timoshenko, S., and J. N. Goodier, Theory of Elas- ticity, 2d ed., McGraw—Hill Book Co., New York, 1951. Ward, I. M., and E. T. Onat, "Non-linear Mechanical Behavior of Oriented Polypropylene," Jour. Mech. Phys. Solids, 11 (1963), 217-29. APPENDIX I ANALYSIS OF STRESS DISTRIBUTION IN TAPERED MODEL As stated in Chapter II, any model for which there is a known analytical solution for the stress distribution can be used for calibration. Such a solution for a wedge loaded with a concentrated force at the apex is given below. Combining this with experimental evidence obtained for a material with known properties, the value of radial stress along the centerline of the tapered model used throughout this study is obtained. Timoshenko and Goodier [68] obtained the solution by assuming stresses which satisfy the boundary conditions. An unknown constant is evaluated from equilibrium con- siderations. Frocht [18] used the Airy stress function m z Cre sin 8 (Al.l) to obtain the stresses 0']: .. 2c 233—9, (111.2) Ge :3 O, (A103) 151 152 where 9 is measured from the centerline and r from the apex as shown in Figure Al.l. The boundary conditions are satisfied by such stresses, and it may easily be verified that the stress function satisfies the necessary compati- bility equation. Fig. Al.l—-Sketch of wedge with concentrated load at the apex. The constant C is evaluated by considering the equilibrium of a portion of the wedge bounded by the 153 straight sides and the surface r = constant. The constant C is thus found to be: _ P C - d(2a + sin 2&7’ (A1°5) where d is the thickness of the model. Combining Equations (A1.2) and (Al.5) yields: 0 = 2P cos 0 (Al.6) r rd(2a + sin 2d)’ The equation relating order of isochromatic and stress was given in Chapter II as: Cod n = T (01 - 02). (A1.7) In the case of the wedge, 01 = Or and 02 = 0, so C _ o 2P cos 9 n _ T—' r(2a + sin 2a) ' (Al'8) Equation (A1.8) shows that on the centerline of the model, when 9 = 0, the isochromatic fringe of a given order is circular. For the model used in this study a = 7.12°. Since cos 7.l2° = 0.9922, the fringes may be regarded as circular for the entire wedge portion of the model used. 154 It is interesting that the solution presented above, through Equation (A1.8), does not restrict the value of the angle a to being small. Hence the solution applies as well to a semi—infinite plate, where a =-g. For the modified wedge used in this study (soon to be described), the small part of the wedge is transformed to a section of uniform width. Hence the apex of the wedge is not on a boundary. The radius r, therefore, is not easily measured. However, the width of the model can be measured with accuracy and speed. This can be used to determine the nominal stress which can in turn be related to the radial stress or as follows: Onom =-———E——— from Figure A1.1, d(AlA2) B = r cos 9, AlB = B tan a = A1A2/2, AlA2 = 2'_§ tan a = 2r cos 9 tan a, o = P (A1 9) nom 2rd cos 6 tan a' ° Let r nom where k is a stress concentration factor. From Equations (Al.6), (Al.9), and (Al.10), 155 2 cos 9 (k) 1 2a + sin 2a = 2 cos 9 tan d’ from which 4 cos2 9 tan a 2d + sin 2a k = . (Al.ll) Along the centerline of the model (wedge), where the fringe order for birefringence or moire patterns is read, 6 = 0 and _ 4 tan a (k)8=0 _ 2a + sin 2a (Al°12) For models used, a = 7.13°. The value of k along the centerline becomes: (k)e=0 = 1.011. (Al.13) Thus r nom The above analysis is valid only in that portion of the wedge—shaped portion of the specimen where no disturbance in the stress distribution exists. The expression for or shows that or —+- m as r —a’ 0; hence it is necessary to modify the wedge shape in such a way that the smallest cross—section is capable of carrying a load sufficient to cause a meaningful stress in the largest portion of the wedge. This is accomplished by transforming the smaller portion of the wedge into a prismatic "shank," as shown in 156 Figure Al.2. This shank is long enough to provide a uniform axial stress over a length sufficient to accurately measure order of isochromatic, longitudinal strain, and transverse strain. On the end of the prismatic portion opposite the wedge, smooth transition to a corner is made, thus providing a point of zero stress where the zero-order isochromatic is found. The small holes near each end were used only to align and hold the model for machining, as discussed in Chapter II. Transition region Sh7nk { t— % s44rWedge :: . aJ -..--- :_f [Sr—J: 0. 40" beginning of fillet L 1 1 1 V I T 2" 2.1" 11" Fig. Al.2—-Sketch showing shape and approximate dimension of tapered model used. Since the region of transition from the wedge to the shank does not lend itself to mathematical analysis for the stress distribution, an experimental method is used. The stress concentration factor k, defined by Equation (Al.10), is determined continuously from the wedge portion, where its value is 1.011, through the transition region and into the shank portion, where its value is 1.00. This 157 experimental determination is easily made with a material known to have a momentarily linear stress—birefringence relation for stresses below those encountered in the shank during the determination. Polyester Resin Palatal P6-K in the cured state was known from previous tests1 to possess such properties and was therefore selected. Because the stress-birefringence relation is linear, each integral-order isochromatic fringe represents a given difference in principal stresses. But since the distribu- tion is radial in the wedge, this given difference is simply an increment of the radial stress or. On a picture of the isochromatic fringe pattern, taken with a constant load on the model, the location of each integral fringe was determined on the centerline. In the wedge portion, where k is known to be 1.011, the radial stress at each fringe location is computed from the nominal stress. Now, when the increment of or represented by each integral isochroma— tic fringe is known, the value of the apparent or can be determined at each point on the centerline where a fringe crosses, through the transition region and into the shank. At each of these points the nominal stress is computed. Now the stress concentration factor k is determined at the 1Since both stress distribution and isochromatics are radial in the wedge portion of the model (Fig. Al.2), the momentary linearity of the stress-birefringence rela- tion can be checked up to the maximum stress in the wedge portion. This stress should not be exceeded in the shank during the test for determination of the stress concentra- tion factor. 158 discrete points where fringes cross the centerline. The known value of k in the shank affords a check. The value of k may now be plotted against the distance along the centerline. The plot used throughout the photoelastic calibration studies is shown in Figure Al.3. The beginning of the fillet, located by means of a micrometer caliper, served as a reference point. Two scales are shown on the abscissa because in scribing the scale on some models metric units were employed; on others the scale was in inches. 1.024 r nom 1001‘ . ' . o in from beginning of fillet 1.00 ~\ \ \ 99. \\ cm from beginning ' \ of fillet \ \ .98- \ \ .97- Fig. Al.3-—Plot showing variation of the stress concentration factor k along the centerline of the photoelastic model. Additional points on the plot may be obtained by considering half—order fringes or by changing the load on the model. 159 The external load at the large end of the taper disrupts the radial distribution of stress as given by Equation (Al.6) near the end. The width there is approxi- mately 2.5 inches. The isochromatic fringes are very broad in the largest portion of the taper, due to the small stress gradient, making fringe location difficult. Hence no readings were taken in the portion 3—4 inches from the large end. The lowest useful stress was then approximately twenty percent of that in the shank. A similar calibration could be performed for the strain distribution. It is more complicated because an integral moire fringe does not represent a particular strain but only a certain displacement relative to another fringe. Furthermore, the effect on the moire pattern is not as pronounced as on the isochromatic pattern for the following reason. The isochromatic order is related to the difference in principal stresses (0 — 02) or, in this 1 case, (or — 09). Hence any small transverse stress 09 affects the isochromatic order directly. The strain in the radial direction would, on the other hand, be affected by a transverse stress component only through the Poisson effect. Since Poisson's ratio cannot exceed the value of 0.5 (in an elastic material), the effect of such a trans- verse stress component on axial strain is less pronounced than upon the isochromatic fringe pattern. 160 Poisson's ratio is less than 0.4 for the materials tested in this study. Irregularities of the moire fringe pattern in the transition region were not apparent. On the mechanical models the stress concentration factor was therefore assumed 1.00 in the shank and up to the point of crossing in Figure A1.3, and 1.011 for the remainder of the model. Typical isochromatic and moire fringe patterns are shown in Figure Al.4. F'" «A 5 :5 mum l :2 All! m , I (a) (b) 7! mun» ) Fig. Al.4——Photographs of (a) isochromatic fringe pattern and (b) moire fringe pattern. APPENDIX II TECHNIQUES OF MOIRE STRAIN ANALYSIS Difficulties Presented by a "Conventional" Method In order to plot strain as a function of stress, at a given time after loading, including not only the linear but also a portion of the nonlinear region, at least eight points seem necessary (see Figure 3.5). Though the general shape of the stress—strain curve can be established with only a few points, the accurate evaluation of the linear limit stress value at a specified deviation from linearity of strain requires eight or more. These considerations led to the attempt at measuring the average strain over a 4—inch gage length by means of a mechanical extensometer on eight tension specimens of uniform width. The extensometer consists of an invar steel bar attached at one of its ends to the specimen, with the bar resting lightly in a guide cemented to the specimen near the other end. The displacement of a fine line scratched on the bar, 4 inches from the attached end, rela— tive to a similar line on the guide, was measured with a telemicrosc0pe supplied by Gaertner Scientific Corporation. 161 162 This method promised three distinct advantages: (1) high sensitivity (better than 1 x 10"4 in/in); (2) no adverse effects, such as stiffening or indentation of the specimen; and (3) repeated use of a single extensometer. Inherent in this method are some disadvantages. Only one specimen at a time can be tested without special load frames, designed to accommodate multiple specimens, and provision for readily and accurately moving the tele- microscope to obtain successive readings on all specimens is also necessary. A load frame holding nine specimens was used by Marin and Griffith [38], with electrical resistance gages measuring strain. The expense of building such a load frame, with the additional telemicroscope carriage, was beyond the means of the present study. The series of eight or more separate tests require at least eighty days for obtaining the creep data desired in this study; nevertheless, it was attempted. Eight specimens of Polyester Resin PS—l, with a prismatic portion exceeding 5.2 inches in length and a width of 0.40 inch, were cut from a plate approximately 0.130 inch thick. From the data obtained by separate tests at different stress levels curves of strain vs. time were drawn for each stress level. Six tests at successively higher stress levels, extending well into the nonlinear region, were followed by two at stress levels by then known to be near the linear limit stress. Although it was difficult to repeat readings 163 closely, little scatter was evident in the data. From this plot data was deduced for strain vs. stress at selected times. In this strain—stress plot glaring discrepancies appeared. The eight points available for plotting each curve were scattered excessively for purposes of accurately determining the linear limit stress. The points corre— sponding to the last two tests appeared four to five per— cent above the curve established by the other six points. Several causes, acting alone or in combination, might be responsible for such erratic behavior. As men— tioned in Chapter II, when more than one specimen is used in a calibration test, they must all be machined and handled identically. For convenience and consistency all eight specimens were machined at once. But then how are they to be handled identically when they must be stored for various periods of time awaiting testing? Since the time- edge effect was known to be significant on some plastics, the machined surfaces were coated with silicone grease and the specimens were stored in tightly closed polyethylene bags. Perhaps this precaution was not necessary-—or per— haps it was necessary but not effective. Possibly it even caused a degradation of the data. Lack of homogeneity could also cause the discrep— ancies. Frocht and Thomson [23] found variations of up to seven percent in birefringence, in eight different models 164 cut from a single sheet of celluloid. Some variation in mechanical properties seems just as likely. Reloading for shorter time periods produced roughly the same magnitudes of strain response but a different rate of creep, particularly at the highest stresses encountered. No data of value to this study was obtained. Suitability of Moire Method Because of the time required by further attempts to use the method discussed above, it was abandoned in favor of one employing the moire method on a tapered specimen with a shape identical to that used in the optical test. Its use on such a model yields strain data for any (every) value of stress within a certain range, this range quite naturally being about the same as for the test of optical properties on the same material. Hence one model yields all the stress—strain data required by this study. The theory and methods of interpretation of moire patterns are well—established [15,43,51,55,56]. Theocaris [66] recently reported on past efforts and presented an extensive bibliography on the moire method. Low sensitivity has generally limited the usefulness of the method to the measurement of the relatively large strains. With two initially identical line arrays of density 500 lines/inch each fringe, formed by superposition of the two arrays after one is deformed, represents a rela- tive displacement of 0.002 inch. By counting half fringes 165 a relative displacement of 0.001 inch can be detected. Since strain in a given direction is given by the displace— ment gradient in that direction, fringes must be closely spaced in regions where the gradient is small or changes rapidly if serious errors are to be avoided. Fortunately, no such rapid changes were present in the tapered specimen used for this study. Accuracy of the moire method can be improved by many techniques [9]: (l) Higher—density (i.e., smaller—pitch) arrays produce more fringes with consequent increase in precision. (2) Variable~pitch arrays allow compensation- type determinations of fractional fringe orders, thus increasing accuracy. This is essentially the same as using an initial mismatch of pitches, producing an initial pat— tern, and subsequently yielding a larger number of fringes. (3) Photocells can be used to measure light intensity at a point. The intensity, in turn, can be related [53] to the maximum and minimum intensities corresponding to integral and half—order fringes, thereby yielding a continuous rela— tionship between light intensities and displacements, and hence strains. Sciammarella and associates [52] obtained better results by this method using a 300 line/inch grid, 6 inch/inch, than by yielding minimum strains of 20 x 10— the conventional method using 1,000 lines/inch in a pre- vious test under similar conditions. (4) Various methods of fringe interpolation have been successfully employed 166 [54,58]. Sciammarella and Lurowist [57] obtained partial fringes by observing the normal moire pattern in a field of polarized light, thus taking advantage of the diffraction characteristics of line arrays. Low [36] achieved the same result using diffraction gratings instead of ordinary line arrays. These gratings promise several advantages other than the one mentioned. None of these refinements seemed necessary in this study. The investment in time and auxiliary equipment required to utilize them seemed unwarranted. As mentioned above, no rapid changes in the displacement gradient occurred. It was seen in Chapter III that a 500 line/inch array yielded an adequate number of points for plotting a smooth curve of fringe order vs. length along the specimen. Loss of accuracy resulted mostly from evaluating the slope of this line rather than in constructing it, and this evaluation might be improved by means of a suitable data—' smoothing technique if necessary. The choice of 500 lines/inch as the density for model and analyzer was somewhat arbitrary. The pitch most appropriate for a given application can be specified only within some range. It should be small enough to yield sufficient response for the smallest strains anticipated but should not be so small as to produce so many fringes in regions of highest strain that they can no longer be clearly distinguished. The difficulty and expense inherent 167 in producing fine arrays will usually dictate that the largest permissible pitch be used. Production of Line Arrays-- General Methods Methods of producing line arrays vary widely, depending upon such factors as the density of the array, type of model material used, equipment available, and the number of arrays required. Reference arrays in various densities are available on a stable transparent backing, such as glass, with cost depending upon the density and over—all size. Duplicates of these can be made by contact printing. Suitable arrays can also be made by photograph- ing larger, inexpensive, coarse screens. Line arrays on the model are much more difficult to produce. Of course, they must be accurate in line width and spacing and must follow the model perfectly during deformation yet offer no significant resistance to deforma— tion. Model arrays are usually produced in one of five ways: (1) The array is scribed or indented into the surface. (2) An array is photographed or contact printed on stripping film, and a very thin film carrying the emul- sion is cemented to the specimen. (3) The array is printed, by contact, in a layer of light—sensitive emulsion coating the specimen. (4) Commercially available gages with arrays etched in thin metal may be cemented to the specimen. When the cement has cured, a metal backing is 168 stripped off, leaving only a series of thin metal filaments attached to the specimen. (5) A diffraction grating is "printed" by contact in a layer of liquid resin, coating the specimen. The problem of choosing the method most suitable for a particular application has many aspects, some of which were mentioned at the beginning of this section. Applica— tion of an array on plastics presents problems not present for metals: chemicals used for etching may attack the plastics, scribing may introduce stress concentrations or fail to yield sufficient contrast for observable fringes, and the heating required for the cure of some coatings is excessive for plastics. The first method mentioned above——namely, scribing or indenting the array into the surface—-can hardly be considered for plastics. It has, however, been used quite successfully on other materials. Using the mechanism from a rotary microtome, Douglas, Akkoc, and Pugh [14] obtained consistently successful gratings of line densities to 12,700 lines/inch on flat metal specimens. Bell [2] obtained rulings in the form of cylindrical threads with densities as high as 30,720 lines/inch using a specially modified lathe. The use of stripping film, as mentioned secondly above, offers the advantage of using directly an array on film, thus eliminating the additional step of contact 169 printing on the emulsion-covered specimen as required by (3). Since stripping film was used extensively in this study, its use will be further discussed in the following section. Before a suitable cement for bonding the stripping film to the specimen was found, the third method was tried. In this method the specimen is coated with a light— sensitive emulsion which is, in turn, exposed by contact printing to the desired array on glass or film. Because most of the available emulsions are relatively insensitive and because intimate contact is required, a vacuum printing frame and carbon—arc lamp are usually used in the opera- tion. These were made available for use in this study by two local firms——one a lithographer, the other a photo— engraver. Although this method is probably the most widely used for applying arrays to models, success was not achieved in the course of this study. The greatest diffi— culty was the lack of an inexpensive array on glass or film which transmitted the light exposing the emulsion with uni- form intensity. Glass plates with the desired 500 line/inch array scratched into the surface were known to be available. Their cost was prohibitive ($16 per square inch from one manufacturer) for the size required in this study (approxi- mately lO-inch length). Film reproductions of these arrays, 170 which were later purchased, were not known to be available at the time the method was attempted. An array of 500 lines/inch was, however, obtained on film by photographing a coarse, inexpensive screen of 65 lines/inch. The lack of uniform light transmission by the screen caused a corresponding nonuniformity in the film, which was further magnified in the array printed on the specimen. The exposure of the emulsion coating the specimen ranged from too low, so that the emulsion was removed in developing, to too high, so that the lines could not be distinguished. The images on film, however, were sufficiently uniform in intensity for the production of high quality and high contrast moire fringes when two such films were placed in contact. Since these images could successfully be produced on stripping film, the method employing such film directly was again pursued and finally used successfully. Before discussing the technique used in applying the stripping film, the last two methods men— tioned above——i.e., use of commercially available, bondable gages and diffraction gratings——will be briefly discussed. Bondable gages promise the advantages of easy application, lack of stiffening since the grid is adver- tised as consisting of a series of unconnected "dots," and extremely high contrast since the array is formed in metal which is transferred to the specimen. Their chief dis- advantages would seem to be relatively high cost and 171 limited size (e.g., at present the largest available 500 line/inch grid is 3.6 in x 3.6 in at $55). Application in this study, where an area of 0.40 inch by 10 inches was needed, required cutting the largest available grid into strips and joining three strips end to end. In view of the promised advantages, one of the bondable gages described above was purchased and bonded, in strips, to two specimens, one of which was to serve as the analyzer to measure the strain response of the other. After testing it was discovered that instead of a series of square dots the grid actually consisted of a series of crossed lines, thus aborting some of its greatest potential advantages——those of offering no resistance to deformation and of following arbitrarily large strains. Though the slight stiffening was not believed significant, no further use was made of the bondable gages, since cross-lined metal screens of the same density were found to be available from another manufacturer in an ll—inch by ll—inch size for approximately the same cost as quoted above for the bond— able gage of one—ninth the area. One of the metal screens was purchased to make contact prints on film of the cross— lined grid, but it was not used directly as a gage. The diffraction—grating method used by Low [36] offers at least three distinct advantages: ease of appli— cation, high density, and inherent fringe-interpolation potential. The grating is available as a series of 172 "grooves" in hardened gelatin on a glass backing plate. The specimen to which the grating is to be applied is coated with a liquid resin into which the grooves are pressed and held until the resin "sets" or cures. The master grating is then removed and can be used again. Gratings with 5,000 rulings per inch have been successfully employed; 2,500 rulings per inch are more common and generally adequate. With the proper optical system the fundamental moire fringes are reduced to fine sharp lines, thus facilitating accurate location. But in addition, subsidiary fringes can be produced which are fractional—order fringes. These are useful for interpolating displacements within the funda— mental period, an important factor for very small strains or high—strain gradients. Although the increased accuracy of such a method was not required in the present study, the low cost and appar- ent ease of grid production might have dictated its use. Unfortunately, information about the method came too late for inclusion in this study. Grid Application Method Employed Of the five previously mentioned methods of applying arrays to the specimen and analyzer, three were tried but only one proved successful-—namely, the one employing stripping film. In this method the desired array is 173 imprinted in the emulsion of the film, either by photog- raphy or by contact printing. A coarse screen of 65 lines/inch was photographically reduced to 500 lines/inch on stripping film in the first attempts of this study. Later the commercial availability of a 500 line/inch film reproduction was recognized, and one was purchased. It was used to make contact prints on stripping film. These prints were more uniform but gave slightly less contrast than the photographs. The stripping layer or membrane carrying the emul- sion is quite stable, considering its thickness (between 0.0004 and 0.0005 in). This stability is necessary for the usual transfer process of stripping (wet or dry) and cementing to a new base. While its stiffness is not sig— nificant in stiffening the specimens to which it was applied, it is also inadequate for allowing one inexperi- enced in its handling to transfer, after stripping, a piece 0.40 inch by 10 inches to a specimen without extensive distortion! If, however, a cement could be found giving adequate adhesion to allow the backing of the stripping film to be peeled off after cementing, then the problem of distortion would be drastically reduced. The stability of the membrane creates further demands upon the cement, since no creep of the cement is allowed if the bonded membrane is to follow the specimen upon straining. 174 Eastman 910 cement was found adequate for the task. Its "instant" curing adds another advantage. A liquid cement of the same material as the specimen has been used [12]. Besides the lack of availability in liquid form of some materials, the excessive curing time of most materials handicaps their use as a cement. The procedure used in cementing the thin strip to the specimen was developed without the advantages of prior experience or any instructions. The specimen to which the membrane was to be bonded was thoroughly "degreased" and cleaned with powdered abrasive. The strip of film was cleaned with a cloth moistened with a very mild acetic acid solution. The film, both membrane and backing, was then positioned and one side taped. The exposed part of the specimen surface, as well as the edges and the film back— ing, was then coated with a releasing agent to prevent the "squeezed out" cement from adhering to the specimen. Care was necessary to avoid getting the releasing agent under the film. The film was then raised with the taped edge acting as a hinge. The catalyst was applied to the film and a layer of the cement was applied to the specimen near the taped edge of the film. The specimen, film side down, was then placed on a surface plate, and uniform pressure, via weight on a straight bar, was applied from the back (top) of the specimen. With high pressure applied for a few 175 seconds the cement cured. Excess cement was then wiped off, along with the releasing agent. The backing was stripped from the membrane making the specimen, with line array attached, ready for testing. Other details of the testing procedure are given in Chapter II, Part B. APPENDIX III DATA FOR OTHER MATERIALS TESTED Polyester Resin PS-l 176 n Fringe Ordor, 20 IO I I 177 / Polyester Rosin PS-l / X=546 nm, d=0.l30in. / I I l I I I J J 3,000 4,000 |,OOO 2,000 0' ( psi ) Fig. A3.1-~Isochromatic fringe order n vs. stress, t = tl,t2, . . . ,tn for PS—l. I I I I I I T I I 178 oHI-mm “OW Qbo o o o aNbo-Hb .m> c umoHo mocfium UflmeoncuomHllm.m< .mfim Anna—05V 057—. n o .mEHp oo~ om. ON. om o¢ o . 4 . . . . 4 7 . q a . . . .J . . O .:_on..ouu .Ecm¢mu4 - Tma Emma 532:0; 1 -¢ 3a 000.. u b .... 00¢.N a .0 .2. 000.» u .0 ..a 000.» u .0 .... cow; .b .... com... a b 33 000.0 a .0 :9 nth.» a .0 n n p p p n n p u ‘ 19pm ebugad I I , .Hnmm now an. . . . .mp.au n p .mmmuyw .m> coapmmcon mmmpcmuummllm.m< .mHm A in 0 b 000.0 000.0 000. v 000.n 000. N 000. _ 0 1 a . _ d d 4 d _ d d _ a a a 4 4 A a d r \\\\\\\ .oNo ...mm Emma .Bmorod T Lamas T \\ I 1 .960 9 T N ... 1 .00; T . . 1 00.. I .0900 no. 03an I; o . 2... 1 ..00 .00 :On O._A:b VW n... x . I .oo IOVN W*C_OQ 0‘00 II 0 LOO.N , AnNN uogwbumg abowoomd 180 0¢N Q a .Humm mom o . . . «anHD u b .mEHp .m> COHymmcon mmm¢cmuummllw.m¢ .mfim Om. _ 3.30; V 2:: 0N. _ 00 O _lwm J— Emom .2335 .3 com; u b b 3n 00¢.N b .2. coo.» b 3a 000.» b L .3 com... b 3n Coot—v b 3n OON.m :3 000.0 I b r _ _ O 10.0 L L. r0._ 10.N uogwbu0|3 abosuaomd 181 fi I fi’ I I I I T 1 I I I I I I I I I 20- I8 " O __ Ih 001 b n -- 240h 00' IS » - I4» I2 ~ IO ~ Fringe Order , n (D r r. 6 r / 4~ / , -' Polyester Resm PS-l X=546nm, d=0.l30in. 0 1 1 1 1 1 1 1 1 J 1 1 1 1 0 0.2 0.4 0.6 0.8 |.0 |.2 L4 |.6 l.8 Percentage Elongation Fig. A3.S—-Isochromatic fringe order n vs. percentage elongation, t = tl’t2’ . . . ,tn for PS—l. I I 182 Epogy Resin PS—2 183 oNImm “0% ago 0 o o oN¥aH 2.3.0 p u v .mmmupm .w> c umouo mmcwuw UHmeOHQUOmHIIm.m< .mfim 0000 OOOV 000m OOON # q a . T d1 n q d dl u q - ...: _N_.0... u .65 2.0 "4 Nlma £25. 383 / \. /[ ....o.oo..o ’\ ,\/Ill..oo.oo..o /I ...oo .oo ... rI..oo.oo..~. / ,l..oo.oo._on /||||I .00 :00. .oolgoew h b I L F b P h h h h b b N. t. u ‘ uplo obuug 0¢N CNN com 00. o C o NImm HO.“ O 0 0 0 .mo.Ho n o .osflp .m> c noouo mmcflum UH0MEOM£UOmHIIb.m< .GHm 00. 352: «EC. 03 ON. 00— 00 0mU 0¢ ON 0 184 d u u u la .5 _~_.o no 6:. mam . x Numa Emom 03.5 .2. 000.N u.0 .3 oomd ... b 3.. 000.0 ......0 ..a 000.» u .0 u q _ a a .. d . .... coo... u b 2.. oom.¢ ... .0 2a coo...“ u b 7.. 000.0 a 0 3.. 000.0 u .0 obuud u ‘ JOpJQ 0. N. .3 185 ONImm “0% up. 0 o 0 “NpaHu. up .mmmu#m .m> cowummcon manpcwuummlIw.m< .mflm 2»... b 000.0 000.0 000..v 000.0 000.N 000.. 0.0amsvaI i O o._.:b.w I I x O //’ / // / ..eo .oo ..o l , oo oo o / . .. . g /. N :9. Emmy. axoam / :00 .00 :N. . /l .00 .oo ..oo .oo ..oo. o .00 go: 0 N0 ¢.0 0.0 ad 0.. N.. ¢._ 0.. uouobuo|3 obosuooud 186 C ONImm ”0% Os 0 o o «N “H I .0.OI $.52: 25... .mgao .m> cowpmmcofim mmwucmuummllm.m< .mflm .... 000.? «.0 ovu 0mm 00m 00. 00. 0v. 0m. 00_ 00 00 0w 0m 0 0 . N I mm 58m .2... No 1 c0 3. ooo.~ u b . .... oon.~ ab II 0.0 . _.. ooo.» «b a no E. o9...» ab . .... ooo... u .o o._ uouobumg abowomad 187 oNImm “OM Syn o o o «N «H p mmmucmuumm .m> c umwuo macaum UHpMEou£UOmHIIQH.m< .mam a u u .GOHpmmcoHo 5.80.3.0 002:3. on. 03w m. ,0. o. N. 0. may 0A0 ego Nay T w 5&3"... .255... I N Iwn. Snow. 08...“. .00 ..0¢N .... n ..oo .oo..~_ .. 4 ....o .oo ..o I- o ofiuug u ‘Jepio 0. N. .1 0. 188 Polyester Resin Palatal P6-K 189 OVmImm HOW Eu.“ 0 o o “mflufinflfl. n” M: ammmuuvm .m> Q vauo mmcflhm UHmeOMLUOmHIIHH.m¢ .mflm :2: b 000.0 000..? 000.N - _ 000.0 000.. I ona 3.2.... 5.2... 3.320.. \ .Enmmouo semen... \ .0. u ‘ JOpJO ebuud 190 oxImm “om Gbo o o o oNbpo " o .oeflp .m> c noono mmcflnm oawmeouwmmMHIINH.m< .oflm .aasozv 0 Foo. oo oo o» oo o... ov on om o. o q. A . . . . . . a d a a J a . q 1 O .sonmo no is: we... u A ._ x Ion. .22.... 53... 3.3.3.. . , a N I it.- A‘ a .3 oom _ "b .n r , . w .... ooo._ u b v I m 1 .1... ooe.N . .o . . o KI . I 1 .2. So.» a b . s K. 1 .... oow.» .. b . o W .... com... u b o .2. 0006 ... b 0., 191 onImm MOM Qua.“ o o o .m .H u. our .mmmupm .m> coapmmcon mmmpsmuuomllma.m< .mflm .oo ..oo. :3... 000.¢ 00N.¢ 000.0 000.0 00¢.N 000.. 00m.. 000 d u dl u d l— d u u. d u d u d T v. Ion. .22.... 5.3. 3.3.3.. .T C T 0 / I , // / I I I z /...oo ..o ..o r // ..00 .00 :0 0.63.01 ..I x I / :00 .8 gm 0.0ANNbvw II o ..00.00..00 «Lion. Boo I.. . N0 v.0 0.0 0.0 0.. ~.. qouobuma obowoonod oVH-Imnm “OM an o o o «thnfl H .0 .mEHM .m> GOHpmmcon mmmucmoummII¢H.m< .mflm 192 7.30... 08.... 00. 00 00 0h 00 00 00 on ON 0. - q 4 q d j q q - q d u - Id 1 o I v. Iwa .3025 Emma 332:2. I I N.0 .... oo~._ u b I #0 .3 ooo._ "b I .3 00¢.N u b 0.0 I .3 ooo.m ....o 1 10 I .2. 000.0 u b I . I0.. .8 ooN ... u .o ...... 000... ... b I 1N; . b . b p b . p p p . . - p P uouobuom abowamad 193 .Klmm How 00. . . . .Nu.ap u p .coa0mmcon oomycounmm .m> c umouo mmcfiuw UHmeOMQUOmHIImH.m< .mflm 5.30.3.0 «02:33.. N.. 0.. 0.0 0.0 ¢.0 N0 1' w v. I0... .22.... Emmy. .3339”. \\ .. . . . 1 .. . £00 114 T. . £0 IIU . . .00 gnu..Io u ‘ Japao bound 0. IIIIIIIIIIIIIIIIIIIIIIIIIIIIIIII IIIIIIIIIIIIIIIIIIIIIIIIIIIIIIIIIIIIIIIIIIIIIIIIIIIII