w t a . v a I V c u, .. . . .V. . ‘ Y. .. u 0.. we .‘t' 3 ac. . ' 9.3.031 .q .. 2‘. av I‘L‘ I-n‘. .. I. Q a Q. -. V“A.P..v 1! 119! v v Q I v. .0 vv ”1.37.. §t.v\fivv%f1¢v v. . tu. c g ’ u 9 ., 7 \ Vibfi. tigrt'vh.~| (-2 . r .v v c q Q! ‘ t a v .ifi‘vov=uv§v‘ewev Eel!!! . .1 fl... V :Q. ‘ A b ~ ‘ .QoNWKOJGUAK. , v ‘Hvtvztxo . 5: u s QR! $frt>ts 1.50. ~ . vs: ¢3~1~731$w , v .. . V .3 r . 1.9!... vfk‘ .o . . ... 1.37mv3... Erhu.‘ u. g .3 . V c c a V V E yvvrvtoQtfiQIX V narysvvl: v: v n. .V r . y. t . e. 9... \A iliut“ V V . : 3!...vv: 21.: Q ¢ r. . . i .. V v +‘1 .\. ibis: I. 1 1.. rotslrclvtvf‘i' 2 y p. u c. r _ “3.5.31 vv‘ 7 .. R3334 «'97 I ~ 4 , ~ he . up.“ 2919 v 286 V V 9 «Riga»... s. . to vv ‘ . or . t a... sit. v . .V . ‘. , a o .kxu... ofiv : u V. 05... 1 viv? 5 YO s1? \ .V o.¢ 9 i \fifihc _ . 917..»Nvfn ub‘z» L. Y9}..01VIYOOI VViafi21. - v . #7720. .97..vuélvvutthsflwIHWLIVViSfifiZ.u v \ YER” k; ‘brvvlvfhk: knurbiv‘V'vag‘LT‘. (in; I'th. ~ v . c n o 1 z 9.: 1. 7r. r‘hu?’ v‘t? V‘fi9avvin.t!..s+5’rvv~.rfi.r C ‘ 0 Q‘.%D‘.f!ig .x. 5 11.“. v." ~A$O)\1!1!2I§viiiili “Ry“... ~ I v :30 ‘2!- u§ Evsfiv fl... 5 (5.1.7.. 6.3.?th furl}: o..~‘v§> 1.. v ‘1'! Vt: ht civfhhiot.» .3» . 1 . u..>-rv-.:~§1YV ‘a VCZ“ tirnifizti u g oh. .5, g. 5. (iii A.\\¢.A .‘3. 9s p 0.. §Z.VX»Itv9.z 3...? .fiL-tiiun‘lvgov .336 ~ . t 1 this c6 :th «if. tutti i c . $.91. I‘VQOVO 0.16? h»! QYCIi 7‘ Y5!!\ QB! i ‘1'. if! ..w.!..v.t..l ittf ‘30, V oi f. uQQ.‘ v uhfltvlitfivvclcv‘v Elfi$v:‘h\hv7t¢a.§ 9.5.0?! 15:}... ~$Iit§v§$|§b~6fi 7.7:... ..t. ‘ti-s‘)?1 V??? izl“l77!?t\’h1!hfll"|~uv~ts v. . OE é}; it vxtl‘?f:c.ot¢l.9. '3‘... I1, vf¥vf t 7!: ti .3"... t t 003's £1 . a. v.0 7.x ..... .11.»..50‘u x‘,...?:t;l§ 15.... c-s‘g97v4uwuu6 .iflflltq.£biui otvqofizvbt & g Eivkfltt‘s \I...%H\A‘w0isdn.l I .3: I.1t£.v:.~v¥lfll1t-%rv.t&fllso§% {fluvialliwnabuziuf.....nvt£.:.1nvlo.V ‘ . txi....v+r.vhv~tflnw5§~ tertalg fizifihnfiibbfbn ..I.I..Z.L....\is.- V11. Ina...“ ilot .\ laid-‘KVIV :L tifig. flit .Aflh« :‘Ei‘ieflflr ‘. kuifi. 0 .‘Efli. 7.1L.“ .«wmnuvvostssu 1: .Q. t‘Tsflflsfifuit} v sittizlov» .2 3: Eu $5“ka ,0}st #1:} t éu‘foi fit it?“ Nt‘UnNNMOVthfl‘ ‘0“ 1 Ah“... “09.1.61 it!!! tt..~ . .erL\.tv ii, 5.... {A (V . i 1%“ \i\‘..t1h.\.hvfirtkhtni.dl{tgefl _ . .vHflLhA .kttii u...09s0§.|‘:.1r¢b .Q 10‘... . (V1 §.A:il£v§o .{o‘g . Qutd‘t... iii.% 75.. 'll «t‘ODOE bititkn; S. vfisfltt? 612.255.... , i.-1fs\..1.u\. .fifi‘mwmt..wi 3 £0; Qttt‘rfrt 3‘): Math». 1?... 2“. ‘79.; .V V. _ tar: . 2.1318 ...1 has... RH“! 0110.... o. ‘1! N...“ fl“... is.“ u . VVQc'VV t! 1' . “a vdvyvflvhubflnvflu ivurruo 1 $57!, .0 :V It... all Qvlmvvt. I V 5-.....» Ptvw... . . Lifi3s..%nlhvnnwfixnu.!1 . . . a!» .. V. O a v 1.»! . 5...? 2.3.1“...thon . 2.31%... t . . 1! VI :2. u . g 1511. vvvvvv L. I ZSQ,‘ | .- . .v. if...» .Ewgbfl..ufi. . u E .. . ‘ '. . . ’ . V ‘ V I V \ 7 In . . . I. . A , fig“ V 5‘... ‘Z&.\; x V fizz..vfllu.§ V . ta? 1 h u _n\1... 5391...}; inw‘l ur “Hui law». innufiubptylfifzyf‘ .1211... aziilnw‘tzb E... I ““11“.”qu I: ' }zz06“£§ 7‘ IIV rdx‘lg,‘ Fur Longtikuwtnl. vi. 1 . rttsbfi $3.3! ‘1‘): V . Vi!!! , iii I t: ifim’ . 11.11.32. #1 Vt: -L.z«urwrm§dr Figs”? , "v‘lell‘h "h'i 3““! {Vi} ‘ifii‘flz 1V V... 5... W02 313.11: {93:307. NW... u.:.11...hdnnwzsfififli{!at3¢!v. .rJum . v"... inv‘i‘n.‘ $ iavv. iv: 13.! .. {but 4 (I! 15!? with—viii: “MTnn Erflfiwhfifi... 5......«3. z . .1VIkyutvv... 5%.... : \‘JHPVVVEGN. vuhflkrgfifififlvvflflwvei" “file... . . it. . 546‘s" Natl—1.x; 19"!!!“ Khuhllrl‘vfigiz!vtévifi V‘szh.v0.‘rst§.ifp$ ‘. ‘lfilr... ‘3‘}; .232 . V v 1.?!» VVV‘Q.‘ lvvfli ‘41)..1'1“ IVCYIILLOAQBUn-vvxh gal. all-Oi? a. 1.3? . . iii“tuhfidfldflvfl%flvfihfi Y... htvblnhvttfin‘ilio Eliggi V 2 .. ! tfihhfil $16... .5‘..vo.~ir.. . Al. yi..i~.{a.o.!ou. I a NV... 1: Wits... \ . .VO).£‘tv\s.s:.€ nfiivlthv Vt!!! . fiuugkig . . 3! late»... .5 .. u , 211%.: 51. .21; a \ loan. i%fi #953 bill E 9‘!“ at. L: . .I {3.9%.351 st. lilé {fight LI: 5 Eflkn ibé is; lg 1 . . . V , 33‘ :5; (”I 5.1.354 .t. V i‘ if: ‘3 ‘0. ha. la! ‘Llu. , V. v “932‘ . Yv!i‘lv%a Q . O. 3i:1§\.a\\v o\‘ I. VIWV'OLVE" ' l. l}l‘§z‘ ‘.Ni 1 ‘3‘ ‘51.; if}: I Iv . n ‘riilv Iv {tillyinucn If 1.... all... 3t”._u.!v\u\all.l. .flnauuqnnu: Ill... .6 anti U1. NH .13... V .3 Vii}... lyimixfli .175. $14.5 3:15.904! . w‘bkai‘iil‘v“ Lill.\..9.v~tuti ‘1flfiufi «HP-V: :Yfifluvhutfiu‘uh‘o .zQ'. 11.113. .n 23“]: . .92. .L‘Iivy. .1 ,1 '0 u .L.Nhgll&9v)151§zi§i~!'§ xii...- . A a dig!) .A9i‘xvt3tu‘uhizl: V. 56‘ .rn\.§.\ri.uhnflvli1v1riv. .‘JVQV‘ a 6-13.95." «2. 13‘ 6,: \ A V. \ivuvfluahkflfirekflwfifaivlto .t‘kflmv; Didn‘t»? u q a .. w x V a _ . . 1?: 097...? 1551: i ilkitwuih~tfi . \vhb‘ON-Wvfiuhb'ithvivivt .v‘nfl Nfi.)€n.‘1§thfiv%u§b ‘0‘): $rxg>tlvflvrv§vfue 1: Vol. .1192 A- {Stunt}: x. RN“ A . . . gxcii} . ,. 113:; t . sirziifiut-.. V : {in {Aug V. A. . V. I I. r V \ .. .1; x ; . v . m 2.. t , i. ,r .w Us 4..\¥?X!.. . | i v 8 I §\.O‘ ‘ VIQE: . 3.3 V . V .5 .u i . . “5. .EE Q $ - .‘ ‘0‘: ‘9}: ti!!! \ «5‘?! I. «3.5.1» I V .‘Ai‘h‘lu‘i‘t . fittflfilficfittfitaa} irifiu. was. t . . . V ‘ 2% s It L , . 1. 531.} 1.1.3, 3.§I§E§ufi%$fiv\u1 V ‘: K - . . L‘Nt.‘ ~ .a‘l.1t i L» ‘Q 2......» .nn... 1 snub . . . V ,u. .in’i: O: ..‘a’%~ALAI\J/\LA w t ‘5 . {gliuw .V....i...E.L..ts§l11.6\\LL!I..NNB»£LLat»... . V . . . . EELY .t. t: v.15... .. V 1552.}... tun...“ 5.13. b At; ~ . . cl E R 7‘0... . u ...... ’’ tirvzi ~ ...... xv: . .. .1 u :3 . V 3 .V. 1 swuvdflqrniue, 1w: . 9.1.3.9: 1:7... Vite.» Eu‘i .nuxumfimwthébmsflnfly fittfltutffiuv 353k. . xVYPhHWfivatlvHM-Hfix. "¥‘¥I\1 9-11. L [347... EYY. if} .i rLiYKr‘iI 91.11“] V ‘ Eit‘gi . . V I. 1‘ it‘d... I‘vv‘vfivl: 1...: V v .' ' '.. . ‘I ~ -.~ .3 ‘\ 5 a ‘r v' 1 ' A ’5‘?!“ I .§ iii]! 3 - saw... ii.w.i1\..t§:§. vizvtx..- - «MM a... 1. Ghana.“ {yaw £31». 3.1.1.»: 51.3%... . )4... MN. 5.113 LiizY‘23. [ult- u.‘ . . . u i. II . I‘d! .V h0‘v§.;‘s . 1 xiii . e. V {L5}! 111—55. . . . . a2~fivsfii§~§t§§§t€kflt 1! It “lair .rtfibhrvlbnaihwsbtliblt 3“; is . V . ..s.s1L..&4 1.. E‘Wfit‘tg Edna». g.» .1... V . ti: . l . I 0!.» fi.....~..uuwavxszuits $$Mflsi§§vbvo?. f. uzi‘ 1.411.! Obi”?! "‘95.. ‘I‘L’Va it: .5 i 311-. E 1!. 1.1.! Av! . . ’46.... V V)? .. “v.5“ ..c.. . a .105“; 1‘. h! I; . . 1 v \Lt’iihir!lzr«.vi} . v; A . . 2!: Ems. 3.131.». :5}... . V , a} xeroblgfivl‘r‘.“ V i. Q... 6...}. V ‘9 . : a h “‘01.. 1W...“ flying-kl! 686.5 . 5| \. s. lt’drx . Aniksrnxb. that if. i.) vi s. ~ \“JQA‘IQOQS .124 . «.31? V . . inhuman. . I 1.55.5”? .Qbsiufi 5.. JUN»... §i1.§{! I, . y . i}. it: . . "b..i w F‘tlxa‘l ‘3 . . K v: 1&5“ v1 .T‘ 1: 1.... 51... .n A . ans . :1 ..... . .ili‘s‘ikgl a. . V «fitslbtvvk veg 20?.th J. .i . 1.. . I If . vfiflnfiflhi‘fimiflg .UD.‘ V : t .r .L: ”why-.91. $3.11!! ‘u‘fiO‘qhtOI‘tlr‘fiL-Q .V )1»... ssVVlLSHviiut‘lV {Q} it}! x. a . . . (1"; ii. \EE to 9.1 . .V viii; it: 1.14%....“ . . .s . . .wvb ¢ .thfiBDn I . 15“} insulaf . v V... v .1 «(L11 . ‘I\Vr 3.2.323... . Kittkkvitix hulklvovslx V.» l ‘t‘.\ it 1.5}..‘1 ikxntixtfiitln Exit? . fistihli.‘ \‘ I. “Vixfiu. «2?. V Fig $1... ... v I . k 5.. ‘ . V. 9V; A L ‘iulk‘lll u. .V .. . 2 :9: V . Z. x . \ I .v 3 v . , A . s u I A, 2:... a . . n q .t I a» . 1. is. . . . .4 r .. 1.31. w t v V . wordy}... | n?vv~¢v fry nt.v.u.\bx.«t.pu. I‘)HsMUF.I\-|MWF;¥9§:‘Vafi!!0\3vus\cbufl§\ . .V . mitt taunt; nflrlhsysbnfitylitfiigtztgiutttxvtté 11...! .3 V. s V r! K‘s-Nat»). $711.... .1. I. v! 9}}!‘1‘tfl‘0‘z‘! . _ kksuvs. . n .51? giving 5.75;".{1‘91 V ' n . kl. aéigxg. \RH‘EE 9hr: Lhfli‘b‘tvfihfiufll it VUltltc‘A 00111:; 1:01.11 1.5.15}. up. xefxflflrxnt. “Rubia . . it‘ll»: iii-9.! i¥t|§ tkulngtixl BEL. 1'. > :2! L‘ICV li‘lhttlt lilnntttl. rash... . olilethtLt. {.nuliix.i Ett‘knit kl, V\« \L 5%.;th \ ttllsl Lib.“ at , v V“ . . .o. hi. ‘..5 «A... v ‘ K I 5‘37‘ V. .- ~C F. V. 8|..er .) ‘. 0‘91 ’ . 1.13m? . met-0mm gut-e .. Umvea’sity . _i_ ,.._ . ~ v This is to certify that the dissertation entitled THE MELT RHEOLOGY 0F A-B BLOCK COPOLYMERS WITH SPHERICAL MICRODOMAINS presented by Ekong A. Ekong has been accepted towards fulfillment of the requirements for Ph.D. degree in Chemical Engineering é/«s2 {/83 W MSU is an Affirmative Action/Equal Opportunity Institution 0-12771 )V1531,1 RETURNING MATERIALS: Place in book drop to LIBRAniss remove this checkout from All-{Slllt. your record. FINES will be charged if book is returned after the date stamped below. ROLM USE om DP Net CiRlCUmn: THE MELT RHEOLOGY OF A-B BLOCK COPOLYMERS WITH SPHERICAL MICRODOMAINS By Ekong A. Ekong A DISSERTATION Submitted to Michigan State University in partial fulfillment of the requirements for the degree of DOCTOR OF PHILOSOPHY Department of Chemical Engineering 1983 ABSTRACT THE MELT RHEOLOGY OF A-B BLOCK COPOLYMERS WITH SPHERICAL MICRODOMAINS By Ekong A. Ekong A kinetic network model for polymeric melts that contain spherical microdomains is presented and compared with experimental results of poly(styrene-b-butadiene), (Mn = 232,000-10,000, % wt PS = 94.1). A novel form for the segment distribution in the matrix with a constraint at the point of attachment to the domains is developed. Consistent expressions are developed for the rate of creation and destruction as a function of deformation in flows. A key parameter in this development is the degree of repulsion between segments in the interfacial region. Transient and steady stresses are derived for uniaxial extensional flows and compared with an ABS melt data in the litera- ture. At low Hencky strain rates (made dimensionless with a character- istic relaxation time) an apparent yield stress is predicted dependent on the range of repulsion parameter which correlates with the com- position of the rubbery component. Computations were done with this model also to obtain steady and transient stresses in uniaxial shear flows. These predictions were compared with melt rheology data gathered in this work over a Ekong A. Ekong temperature range of 120°C to 175°C. The shear viscosity data above 150°C indicate homopolymer-like behavior; the data at 130°C indicate the presence of a two-phase structure. The dynamic shear viscosity as well as the steady shear viscosity data show trends similar to those reported by Ghijsels and Raadsen (1980) with triblock copolymer melts at low strain rates. The observed stress growth curves show a 1 with a strain stress overshoot at strain rates as low as 0.01 sec- at peak stress of about 0.5. Estimation of parameters in the theory and the sensitivity of predicted stress behavior to different para- meters is discussed. While the theory is able to predict observed low strain rate behavior in steady and dynamic testing, it does not predict an overshoot in stress growth curves at such low strain rates. In the evergreen memory of Kokomma Bassey-Ubong Ekong Archibong Akpan Udoh Ekong and Phillip Akpabio Ekong ii ACKNOWLEDGMENTS I would like to express my gratitude and appreciation to Dr. Krishnamurthy Jayaraman for his guidance and direction throughout the course of this work. My special thanks also go to Dr. Eric A. Grulke, Professor Donald K. Anderson, and Professor Dennis Heldman for their many helpful suggestions and guidance. Dr. Lu Ho Tung of Dow Chemical Company, Midland, Michigan, for the preparation and characterization of the copolymer sample used in this work. The Department of Chemical Enginnering, Michigan State University, for providing me the financial support in the course of this work. Miss Carole Nicolas for her friendship and assistance in the experiment and data analysis and importantly all my friends, espe- cially Dr. Mohammad Jefri, Dr. Kai Sun, Dr. Chien-Pin Chen, and Mr. Barry Colley, for their towering friendship and moral support. TABLE OF CONTENTS LIST OF TABLES LIST OF FIGURES . NOMENCLATURE . Chapter I. II. III. IV. INTRODUCTION 1.1 Material Background 1.2 Evolution of the Microstructure 1.1.1 Gaussian Network Theory. 1.1.2 Non-Affine Motion Assumption 1.2.3 The Yamamoto Network Theory 1.2.4 The Reptation Theory . BLOCK COPOLYMER MELT PROPERTIES AND THEORIES . 2.1 Previous Rheological Studies 2.2 Optical Studies . . A TRANSIENT NETWORK MODEL FOR POLYMERIC MATERIALS WITH SPHERICAL MICRODOMAINS . . 3.1 Objectives . 3.2 The Rate Terms 3. 3 The Macroscopic Stress Tensor PREDICTED STRESS BEHAVIOR IN EXTENSIONAL FLOWS 4.1 Uniaxial Steady Extensional Flow 4. 2 Results . . 4.2.1 Steady State Stress 4.2.2 Stress Transients . PREDICTED STRESS BEHAVIOR IN SIMPLE SHEAR FLOWS . 5.1 Simple Steady Shear 5.2 Results iv Page vi xi o—J 32 32 32 4o 41 41 42 42 52 52 55 Chapter 5.3 VI. SAMPLE CHARACTERIZATION AND EXPERIMENTAL TECHNIQUES. 010101010101 (33019me Oscillatory Shear Flow Material and Sample Preparation Electron Microscopy Morphology . . The Modified Weissenberg Rheogoniometer . Sample Loading and Temperature Control Rheometric Testing . . . 6.6.1 Oscillatory Shear Experiments 6. 6. 2 Steady Simple Shear Experiments VII. RESULTS AND DISCUSSION \1 \l\l\l 4> wNH Introduction . Phase Transition Temperature . . Viscoelastic Behavior Above the Transition Temperature . Viscoelastic Behavior Below the Transition Temperature . . . 7.4.1 Estimation of Model Parameters . . 7.4.2 EXperimental Evaluation of the Trans- ient Network Model . . 7.4.3 Steady State Predictions 7.4.4 Transient Predictions VIII. CONCLUSION AND RECOMMENDATION . 8.1 8. 2 APPENDICES REFERENCES Conclusion . . Recommendations for Further Study . Page 101 107 116 121 121 125 134 134 136 142 158 Table 2.1 2.2 4.1 6.1 6.2 7.1 7.2 LIST OF TABLES Viscosity of block copolymers vs. homopolymers . 585 samples Model parameter "a" from data of Munstedt Block copolymer characterization Property of glassy continuous phase Phase separated block copolymer melt properties Material constants from experimental data vi Page 19 21 46 75 75 107 118 Figure 1.1 LIST OF FIGURES Variation of block copolymer morphology with compo- sition . . . . . . . . . . . . . . Reduced dynamic viscosity and elastic modulus vs. reduced frequency for SBS 7-43-7 A polymer network with rubber domains . Normalized non-Gaussian distribution functions . Normalized extensional viscosity vs. dimensionless strain rate. Effect of repulsion parameter "a" with e = 0.01 . . . . . Extensional viscosity vs. dimensionless strain rate. Effect of the destruction parameter "c" with a = 0.05 . . Normalized transient extensional viscosity vs. dimensionless time as a function of strain rate. Effect of the destruction parameter "a" with a = 0 Normalized transient viscosity vs. dimensionless time as a function of strain rate. Effect of the repulsion parameter ”a," with P = 0.1 and c = 0.01 Normalized transient viscosity vs. dimensionless time. Effect of the repulsion parameter "a" with T = 1 and e = 0.1 . . . Normalized shear viscosity as a function of dimension- less shear rate. Effect of the destruction parameter, "Elnsa = 0 . . . . . . . . Normalized shear viscosity function and normalized first normal stress difference functions. Effect of repulsion parameter "a," e = .005, E = 0.05 . Normalized shear viscosity and first normal stress difference functions. Effect of the slip factor "5" a = 0.03, e = 0.005 . . . . . . . . Page 26 33 38 44 47 49 50 51 6O 61 6.6a Normalized transient viscosity function. Effect of the repulsion parameter a, e = 0.005, g = 0.05 . Normalized transient viscosity function. Effect of the repulsion parameter "a," y = 3, e = 0.005, E = 0.05 . . . . . . . . . . . . . . . Normalized transient first normal stress difference function a = 0.0, e = 0.005, E = 0.05 . The normalized complex dynamic and steady viscosity as functions of frequency and shear rate respectively w = y . . . . . . . . . Typical EM micrograph of ultra- thin section of poly (styrene- b- butaiene) specimen at x 50,000 TYPical EM micrograph of ultra-thin sections of poly(styrene-b-butadiene) specimen at x 150,000 Weissenberg Rheogoniometer internal (Sangamo Controls Ltd.) . . . . . . . . . . . Cone and plate platen Combined cylindrical and cone and plate platen (Mooney type) . . . . . . . Non-sinusoidal waveform of oscillatory shear stress at strain amplitude of 0.15 . Picture showing test material extruding from gap after a shearing for 8 min; 9 = 0.43 seC'l. Shear insta- bility is due to stress fracture. T = 130°C Quenched sheared materials after a shearing time of 8 min. Left hand specimen sheared at y= 0.096 _1 sec-1. Right hand specimen sheared at y= O. 43 sec T = 130° C . . . . . . . . . Shear stress growth function of S- B at 130° C using the Mooney Platen . . . . . . . Dynamic viscosity and storage modulus vs. frequency of 8-8 at various temperatures . . . . . Complex and steady viscosity of 8-8 at 150°C as a function of frequency and shear rate, respectively viii Page 63 64 65 72 77 78 81 82 82 88 9O 9O 92 96 98 Figure 7.3 7.3a Complex and steady viscosity of S—B at 130°C as a function of frequency and shear rate, respectively Dynamic and steady viscosity of 8-8 at 130°C as a function of frequency and shear rate, respectively Steady shear viscosity function of S-B at 150°C Normalized transient shear stress function of S-B at 150°C . . . Shear stress relaxation of S-B at 150°C The steady shear viscosity as a function of shear rate of 3-8 at 130°C The Steady shear viscosity as a function of shear rate of 3-8 at 124°C Normalized transient shear stress function of 3—8 at 130°C Normalized transient shear stress of S-B at 130°C . Normalized shear stress relaxation as a function of time of S-B at 130°C Normalized shear stress relaxation as a function of time of S-B at 130°C Normalized shear stress relaxation as a function of time of S-B at 124°C Loss modulus and loss factor as functions of frequency at 130°C Evaluation of model parameters using linear visco- elastic functions (dimensionless) . . Evaluation of model parameters using linear visco— elastic functions (dimensionless) . . Comparison of steady shear viscosity with transient network model . Comparison of steady shear viscosity with model Comparison of stress growth results at 130°C with model. . . . . . . . xi Page 99 100 102 104 105 108 109 111 112 113 114 115 119 120 122 123 124 126 Figure Page 7.20 Comparison of stress growth at 130°C with model A = 1.25; a = 0.55, g = 0.05, O = data; Model . 127 7.21 Comparison of stress growth at 130°C with model of A = 4 sec . . . . . . . . . . . . . 128 O 7.22 Comparison of stress growth at 130°C with model of A = 1.25 . . . . . . . . . . . . . 129 O 7.23 Comparison of stress growth at B0°C with model of A = 4 sec . . . . . . 130 O 7.24 Comparison of stress growth at 130°C with model of . . . . . . . . . . 131 A = 1.25 sec . O 7.25 Strain at the stress overshoot in transient shear curves at 130°C . . . . . . . . . . . . . 132 A(B,N) A.. 13 DH 9)? NW (13.0 0 ZIO- A. llm RU (t',t) —+, fin (1 1) Strand distribution function fin: .n (at) - fin/mm (1.2) The distribution function, fin(B’t) is defined such that f}n(R,t) dR if the concentration at time t of strands of complexity i and com- posed of n equivalent random links (of length 1) with ensemble- averaged end-to-end vectors with the range R to B + dB. The term Hn = 3KT/nl2 is the effective Hookean spring constant of an n-link strand, such that HnR can be interpreted as a force on the strand. The terms Lin are the strand creation rates and fi/Ain denote strand destruction rates, with a strand destruction coefficient given as Aih‘ Angular brackets indicate an average value calculated with respect to fin‘ The success of rheological constitutive equations have been mainly based on how well the terms B, Lin and l/Ain approxi— mate the true microstructure dynamics occurring in the polymeric medium. The original ideas on how these terms may be modelled were laid down by Lodge (1954), in deriving the Lodge rubber-like model. In this review it will be useful to state them, in general, and focus on how several researchers have modified these assumptions to achieve useful constitutive equations. Assumption 1: Ensemble-average positions of junctions move affinely and can be identified with particles of the equivalent macroscopic continuum. In particular, if the melt is given a time-dependent homogeneous deforma- tion, we have 3:13 - W (1.3) where 3 denotes an ensemble-average strand end-to-end vector, and V(x,t) denotes the polymer velocity at the place of the position vector x and time, t. The superior dot denotes a time derivative. Assumption 2: At any instant t, the set of network strands in a unit volume may be regarded as mutually exclusive, mutually independent subsets. The probability per unit time that any strand shall leave the network is a function 1/Ain(t) say at t, i, and n. Assumption 3: (i,n) strands are created with spherically symmetric distribution of R vectors, i.e., at a rate which can be expressed as a function Lin(Bat) of i, n,t and the magnitude B alone. Furthermore, at the instant of creation,all (in,) strands have the same distribution as that of a set of free n-Gaussian strands. Using equations (1.1) to (1.3), the constant volume condition (V - V = 0) and Lin(B) expression based on the Gaussian chain assump- tion, a general constitutive equation may be written of the form: t P(t) = im(t,t')B(tt')dt‘ (1.4) ~ ~ -CX) where the memory function m(t,t') is given by A ll m(t,t') = kltzLin(t') exp(-(t.dt"|iin(t")) (1.5) 1n and Here B(t,t') is the Finger strain tensor for the kinematic deforma— tion from past time t' to the present time t. If all the creation and loss rates are constant, i.e., all strands have the same complexity, Lodge's "rubberlike liquid model” results. This model predicts a frequency dependent dynamic shear moduli, but fails to show the dependence of steady shear viscosity on the shear rate or a non-zero second normal stress difference. In order to correct these imperfections, several workers, as will be shown in this section, have proposed empirically different choices of the creation and loss rates, but leave intact the assump— tion that the microstructure flows affinely. If creation and loss rates are functions of instantaneous values of strain rate invariants, various equations including those of Meister (1971) and Careau (1972) are obtained. If the creation and loss rates are functions of the instantaneous values 0f.§££§§§ invariants, we obtain the equation of Kaye (1966). These and other related equations have been tabulated elsewhere in a common notation (Lodge, 1974). Most of these equations, usually characterized by many adjustable parameters, predict steady shear viscosity dependency on the shear rate and show a second normal stress difference. How- ever, they fail to reduce to the appropriate constitutive equation of linear Viscoelasticity at low deformation rates. The next integral constitutive equations are the strain- dependent (K-BKZ type) equations in which the memory function includes a scalar function of strain depending on the elapsed time, t' + t as a factor. Recent step-strain data have given compelling evidence for such a "strain/time" factorization (at least in the terminal zone of the relaxation spectrum) (Osaki et al., 1971; Laun, 1978). Out of this class of equations is the Wagner model (Wagner, 1979a; Wagner and Stephenson, 1979b) with a memory function of the form A m(t,t') = KtZLjh(I . t',t), 12(t',tDexp(t'-t)/A. (1.6) J 1( J where E is written as an abbreviation for E2. J 1n In this model,assumption (2) is replaced by two independent mechan- isms for strand loss, one due to thermal motion with constant loss probabilities l/Ag and the other the survivability of strand at the elasped time of deformation denoted by l/Td(t',t). Since thermal motions determine Aj and not td(t',t), then A? and E. would depend on the microstructure of the material, but the J Td would be structure independent. The loss process is thus given by 1 -1 1 73-(tist) - 49 + Td(t'at) (17) J Equation (1.6) is obtained by combining equation (1.7) and (1.5) and by taking 10 tl h(I ,I ) = exp dtlu 1 2 m (1.8) ’1'. The damping function h is chosen empirically to fit stress relaxation data for single-step strain experiments and stress growth data in step-function elongation rate experiments. The resulting h—expression with two adjustable parameters gave a good description of data from a variety of experiments in shear and elongation. A functional of the h-factor was further proposed by Wagner and Stephenson in order to better predict recovery following elongation at constant rates. One major drawback as to the use of two times in A(t',t) in the Wagner model is that it is not in general possible to find an equivalent differential form for the constitutive equation. For some applications, it appears helpful to have a differential equation for the stress tensor. A fairly successful constitutive equation for polymer melts and concentrated polymer solutions proposed by Acierno et al. (1976) expressed the creation and loss processes as functions of structural variables that describe how far the microstructure deviated from equilibrium. This structural variable is governed by an independent kinetic equation of the form d'. % T°——3= —‘.-". P. ’. 1.9 j dt 1 xJ axJ(tr:J/2GJ) ( ) where P5 is the non-equilibrium part of the jth contribution to the extra stress tensor, g given by 11 t Pi = Ej - J m.(t,t')dt'I (1 10) gj is computed from a Maxwellian-type constitutive equation g./G. + T. §%-(= ) = 2T.y (1 11) Interconnection between this model and the fundamental balance law was made clear by Jongschaap (1981) who noted that the segment loss probability function l/Aj in this network model is given by L.1fi/”% -1.” 1 A. T- 25- R. ('12) J J J J Both sides of Equation (1.12) are multiplied by xj. The result is combined with Equation (1.9), xj is replaced by Nj/Njo and the result multiplied by'Njo to obtain dN. N. N. Bil=ii’7i (1-13) J J Here Nj = ij(R,t)d3R is the total concentration of j- segments at time t, and Njo is the equilibrium value of Nj‘ If in Equation (1.13) Nj/Tj is identified with the creation rate Lj(t) = JLj(R,t)d3R, then 12 dNJ. .. N. —=L.-?} (1.14) 1 dt 3 which is the integral of Equation (1.2) over all configuration space. Thus the differential equations forthe structural variable of Acierno et al. are directly related to the fundamental balance law of the Network theory. The Acierno model is seen to allow for the segment creation and loss rates to depend on the deformation through the trace of the non-equilibrium part of the stress tensor. In the context of the Network theory, it is not evident why the particular form of the destruction process was chosen and why it is successful. 1.2.2 Non-Affine Motion Assumption The Network model of Phan-Thien and Tanner (1977) and Phan- Thien (1978) also allow the function creation and loss rates to depend on trgj', but in a more logical manner. More importantly, the Thien and Tanner model altered affine motion assumption of Network theory (see Assumption 1) allowing the network junction to "slip" with respect to an equivalent continuum specified by the macroscopic velocity gradient VV. In so doing, Phan Thien and Tanner introduced an empirical "slip tensor" to describe non-affine motion of the net- work functions and postulated it to be a linear function of the T). rate of deformation tensor 0 E %(VV + VV Consequently, Equation (1.3) is reformulated as R= (W - :2) -B (1.15) 13 in which the parameter E is the slip coefficient. At the same time, Johnson and Segalman (1977) developed a continuum theory of viscoelasticity which allows non-affine deforma- tion. Two deformation histories were defined. One was the deforma— tion history g(t) observed at macroscopic level; the other, E(t), a history of microstructure deformation was allowed to be non-affine with the macroscopic motion. A relationship between these two motions, Xi and E1 at the present time t, in Cartesian coordinates was given by DJ? ‘1)' (1.16) + (‘2‘ xj,i ‘1,j where a is a constant. They then defined a strain measure E(t,t') governed by GE . _ . 5em: ) - {youm ) (1.17) KI ~ EU'J') =1 and substituted this measure of strain into the Lodge network expres- sion to obtain ~ ~ t P(t') = { m(t,t') g(t,t') §(t,t')Tdt' (1.18) 00 As with the Thien and Tanner model, the Johnson and Segalman model predicts a variety of non-linear rheological behavior well, particu- larly, the viscosity is found to decrease with the shear rate. The 14 Phan Thien and Tanner model contains two dimensionless constants e and E that are determined through elongational flow and viscometric flow experiments respectively. For shear flows, Phan Thien showed that the Thien and Tanner model was identical to Johnson and Segal- man's if E = 1-5. The choice of the range of "a (0 < a < 1) as reported by the authors through comparison with experiment was not easily perceived until the work of Lau and Schowalter (1980). They explained the fundamental basis of both models by pointing out that these were objective constitutive equations that can be formulated with a strain measure derived from appropriate linear combinations of the rate of change of material coordinates in the material fixed (corotational) reference and the space—fixed reference (code formational) frames. They chose e1 strain measure related to the combination ¢ expressed in component form as Vj,i e _l 11 A H 1 on v < 1 min "1'71 2 Then a strain tensor (x, t, t') was defined by 3ij (5’ t9t') = 9(59t) E()_(9 tat.) urn I and E(x,t',t') ll IIH If c = (1 - a), the Johnson and Segalman model is obtained while the Thien and Tanner's model results when c = E. Such a rate of deforma- tion measure can also be used to construct anisotropic fluid models 15 associated with dilute solutions (Gordon and Showalter, 1972). A weakness in both models is that they predict damped oscillations in shear stress at large deformation rates. 1.2.3 The Yamamoto Network Theory Yamamoto (1956, 1957, 1959) presented a more fundamental network theory (cf. Lodge's theory) for concentrated polymer solu- tions and melts. The general form of the microstructure dynamics equation (Equation [1.2]) was originally proposed in the first of three papers in which the creation rate function and chain breakage coefficient are functions of the end-to-end distance and orientation of the segments in the flow field. Unlike Lodge's theory, the net- work is considered as non-Gaussian with the result that the free energy of the network segment is a function of the end-to-end dis- tance. Thus Equation (1.1) can be written as E = Z {HN(R)BB 1‘ (B.t)dB (1.19) It is to be noted that the spring modulus HN(R) is allowed to depend on the deformation of the segment so that non-linear springs may be conceived. Yamamoto has shown that physically plausible assumptions about the segment creation rates and loss probabilities lead to vis- cosity that decrease with shear rate, a negative second normal stress coefficient, and an elongational viscosity that first increases with the elongational rate, goes through 11 maximum and then decreases at higher elongation rates. If the destruction coefficient is made 16 independent of the segment extension, the ensuing strain measure in steady elongational experiments is an exponential that increases with time in the orientation of the chains. At a critical rate of strain, the chains are elongated infinitely without breakage leading to an infinite elongational viscosity. Yamamoto then argued that in actual systems, the chains will break down at finite elongations and the destruction coefficient should be a function of the segmental extension. In this lies the germ of ideas behind recent network models which avoid an infinite elongational viscosity by assuming deformation dependent destruction coefficients. Further studies on the Yamamoto theory, especially the non- Gaussian aspect, have been minimal with regard to modelling visco- elastic fluids. Generally, the theory does not give constitutive equations in an explicit form devoid of summations and integrations over molecular variables. However, non-Gaussian network models are receiving increasing attention in the study of rubber elasticity (Chompff, 1977). Recently, Fuller and Leal (1981) have evaluated a form of non-Gaussian distribution function obtained by a Kuth and GrUhn type perturbation of the Gaussian distribution function. They reported no trend in their results different from those of a Gaussian network model. In the present work, a non-Gaussian distribution function will be presented that yields strikingly different predic- tions. The Yamamoto network theory offers clearly a direction in formulating Viscoelastic models of various polymeric systems if an accurate description of its segment distribution function is found. 17 In the Lodgean theory one has no choice but to assume that the Gaussian distribution of the chains prevails. This has been success- ful for homopolymeric melts especially at small deformation rates, confirming the theory that homopolymeric entanglements are a result of weak secondary forces between primary chains, and occupy a length scale of the order of a statistical subunit. This distribution does not represent the microphase structure that determines copolymer melt properties at small deformations. 1.2.4 The Reptation Theory Failure to incorporate molecular variables into the network theory still stands out as one of the major weaknesses of the several versions of the model posed above. Recently, the entanglement con- cept has been viewed in quite a different light by Doi and Edwards (1978a, 1978b, 1978c). The idea that entangled chains rearrange their conformations by reptation, i.e., curvilinear diffusion along their own contours was first introduced by DeGennes (1971). Doi and Edwards have formulated a theory (DE), relating the dynamics of reptating chains to mechanical properties in concentrated polymer liquids. They assumed that reptation would be the dominant motion in a medium of linear long chains. Employing equations from the theory of rubber elasticity, they calculated the contribution of individual chains to the stress following a step strain and related the subsequent relaxation of stress to conformational rearrangement via reptation (1978b). Without further assumptions, notably the ”independent alignment approximation." 1AA, they arrived at a 18 constitutive equation of the BKZ type good for aribtrary deformation histories. Irlparticular, they showed that for monodisperse entangled linear chain polymer liquids, the plateau modulus, zero-shear viscos— ity and steady state recoverable compliance were functions of chain properties as O O GN a M 03 no a M a: a M° (1.20) where M0 is the molecular weight of the primitive chain. These rela- tions agree fairly well with observed data (Graessley, 1980). The only parameters present in this theory are the reptation tube diameter ”a" and a monemeric friction coefficient. Due to the con- straining nature of domains in the block copolymer systems, it is not very evident how the reptation theory can be applied to block copolymer rheology. CHAPTER II BLOCK COPOLYMER MELT PROPERTIES AND THEORIES 2.1 Previous Rheological Studies In this chapter we wish to examine in detail data collected on the melt rehological properties of block copolymers and rubber modi- fied polymers to identify molecular variables affecting their behav- ior. TABLE 2.1.—-Viscosity of block copolymersa vs. homopolymers Polymerb Percent S ViscosityC 80B 0 3.2 65-818-65 13 13 103-538-105 27.5 29 165-528-163 39 118 19S-3lB-193 53 36.5 243-258-245 65 31 33S-18B-33S 8O 28 838 100 5.5 aNote that at 175°C a lot of the domains have been destroyed (Chung and Gale, 1976). bMolecular weights of blocks in thousands CAt shear stress of 2 x 105 dynes/cm2 and a temperature of 175°C (Holden et al., 1969). 19 20 Table 2.1 summarizes the steady shear melt viscosity data at a constant stree, reported by Holden et al. for several different samples of S-B-S triblock copolymer as well as the homopolymers, polystyrene, and polybutadiene with the same order of overall mole- cular weight. It is readily seen that the styrene content affects the melt viscosities of the triblocks. 0n the other hand, it has been shown by Matsuo that the M.W. of the individual blocks affects the morphology of the block copolymer system. Holden explained the large viscosities exhibited by the block copolymers as due to the two phase structure persisting into the melt. Looking at Figure 1.1 we note that randomly distributed cylinders of polystyrene domains in a polybutadiene matrix is the morphological structure of SBS with 39%S content which has an anamolously large viscosity.. Again cylin- drical domains of polybutadiene is the projected morphology for the SBS with 65%5 content, but has a lower viscosity. It can, there— fore, be concluded using Holden's data that viscosity of block copolymer melts is strongly dependent on the morphology of the respective blocks, block length (M.W.) of the thermoplastic block and chemical nature of the center block. Arnold and Meier (1970) presented the dynamic Viscoelastic data for various samples of SBS melts at low frequencies as shown in Table 2.2. We note that the 22-50 sample has an S content of about 35% by weight while the 14-50 sample has about 31%S. They deduced that the difference of the slope d log n'/d log w between the two samples was due to the presence of semicontinous domain phase 21 TABLE 2.2.--SBS Samples SBS Nominal block Slope of log n' Sample mol. wt.a vs log w 10-50 10-50-10 -O.36 14-50 14—50-14 -0.40 22-50 22-50-22 -0.66 14-60 14-60-14 -O.36 14-70 14-70-14 -O.36 MOPS/97b 97 0 aIn thousands bMonodispersed polystyrene, M.W. = 97,000 of polystyrene in the former sample as opposed to "dispersed poly- styrene domains" in the latter case. They further proposed a quali- tative rheological theory for block copolymers system, stating that at very low deformation rates, the molecular network is essentially intact. At intermediate deformation rates, the three-dimensional network will be disrupted and the system behaves as large star- shaped aggregates. Finally, at high deformation rates, these aggre- gates will, in turn, be disrupted and the system will behave as an assemblage of individual non-aggregated molecules. While Arnold and Meier's dynamic data agree fairly with those of Holden et al., it is to be noted that method of sample preparation used in their study, crumbs may have affected the results. Ghijsels and Raadsen have found that the use of crumbs leads to less reproducible results, especially at low deformation rates than the use of compression moulded samples. 22 They also observed that |n*(w)| > n(i)|?=w for all these block copolymers the disparity being greater for block copolymers terminating in polystyrene. They attributed these to the disruption of the domain network structure which must occur in steady flow, but not necessarily in small amplitude oscillations. A further explana- tion of the phenomenon observed above is that the presence of domains in block copolymers disallows some conformation, which would have deen available to chains through entanglement slippage. This, then, tends to increase the elastic free energy of the chains as well as the resulting modulus. The two phase structure can also be manifested in block copolymer solutions depending on the choice of the solvent (Kotaka and White, 1973). When a good solvent for both components is used, triblock and diblock copolymers solutions behave as homopolymeric solutions. When a poor solvent for one component is used, 9.9., SBS or SB in decane, a two-phase structure of insoluble PS in a solution of PB in decane results. The observed rheological behavior is, however, different for triblock and diblock copolymers. In SBS, the PB component dissolved is connected at both ends to the insoluble PS component thus creating a three-dimensional network structure even at a low concentration of the copolymer. In the diblock, there is no formation of a three-dimensional network, but rather a micelle structure in which the PS segments form a rigid core. Upon increas- ing the polymer concentration, the number of such micelles increases and eventually they would be arranged in a regular three-dimensional 23 array. The morphology of such mesormophic structures have been revealed by electron microscopy studies of Gallot (1978). From Kotaka and White's findings, these mesomorphic structures can be classified as elastic gels that can undergo a complete breakdown in structure by continuous shearing. Another strong influence on Viscoelastic properties of block copolymers is the interphase region existing at domain boundaries containing segments of both blocks. Statistical thermodynamic theor- ies of Meier (1974) and Leary and Williams (1973) indicate that the volume fraction occupied by the interphase and, therefore, the degree of compatibility increase with decreasing molecular weight. With increasing temperature, a continuous increase in miscibility would also be anticipated involving growth of the interphase at the expense of the two pure phases, subsequent complete disappearance of the domain phase and then the continuous phase and, ultimately, complete homogeneity. Such predictions have been confirmed experimentally by Chung and Gale (1976) through rheological studies. Using moderate M.W. samples of SBS with spherical polystyrene domains, they noted that at high temperatures, the melt experiences a transition from a multiphase structure to a homogeneous structure. The flow behavior above this temperature is characterized by a Newtonian viscosity at low deformation rates and by low elasticity. Such behavior has been observed also by Kraus (even with high M.W. diblocks and Holden et al. using triblocks). Kraus and Rollman (1976) have predicted the volume fraction of the mixed interlayer for various M.W. triblock copolymer samples, 24 using the theory of Meier. They then correlated dynamical mechani- cal moduli as a function of temperature with the results of Meier. The composition, a of the domain phase segments changes continuously from zero to unity within the range of the interlayer. It was assumed by Meier that the volume fraction of domain phase segments follows a symmetric profile over the interlayer, thus fixing an average composition of the interlayer by domain phase segments at 0.5. This enables one to compute the normalized volume distribution func- tion, V(¢) of domain phase content in the interlayer. The planar * E obtained by applying the principle of volume additivity as: complex moduli, E of the composite for lamellar morphology was * ‘k * 1* _. EE(t) = v8 58(1) + vSES(T) + 61L J 58(1') V(¢)d¢ (2.1) 0 Where VB, Vs and VIL are the volume fractions of pure PB, pure PS, and mixed interlayer respectively; EE, E: are the complex moduli for pure PB and pure PS respectively. Kraus and Rollman, on the other hand, assumed the mole fraction of domain phase segments follows a symmetric profile over the interlayer. They were able to correlate the dynamical mechanical moduli better. Both of these arguments have no factual basis and were formulated for the sake of mathematical convenience. Thus a complete understanding of block c0polymer mechani- cal and rheological behavior will be dependent on the development of a statistical thermodynamic theory for the precise mathematical form of the interlayer composition profile. 25 Gouinlock and Porter (1977) working with SBS samples identical with that of Chung and Gale generated master curves of linear visco- elastic functions using the frequency-temperature superposition prin- ciple as shown in Figure 2.1. Each curve (reduced dynamic viscosity, né and reduced dynamic storage modulus, Gp) has two branches at certain reduced critical frequencies. The low temperature data fall on the upper branches and signify the prevalence of the two phase structure. The high temperature results occur on the lower branch suggesting a homogeneous structure. It is further observed that the critical reduced frequency where branching occurs in G6 data are larger than the critical frequency for UP. It is to be noted, therefore, that modification of the elastic property by domain struc- ture is considerably more pronounced than the effect on dynamic viscosity. Moreover, experiments indicate in contrast to the deforma- tion theory of Meier presented earlier, that domain disruption increases with decreasing frequency. In 'Lufiit of this, the extrac- tion of segments from the domains would be expected to involve long— range configurational rearrangements accompanied by long relaxation times. They then concluded that domain disruption in dynamics measure- ments as in steady state deformation should depend principally on the strain, i.e., strain amplitude, not on frequency, and that it should occur preferentially, if at all, at lower reduced frequencies, where an effect on the dynamic properties attributable to the domain structures as such is alone inferred to exist. Another significance of the results of Gouinlock and Porter is that the relaxation time 26 ‘ Q n T ' 1 l 1 1 1 1 1o 1— ‘ nous - ‘- —1 107 1. _. .. '.. . 1 ° .0. 3 v ‘ °_ it» PM 07 . - g o “\ ‘3 2 1o , 2, 0 00 030. — 10‘ .1 :1 - A {o o 100 31.7 3 r - , ‘5 D 111 0.0 g g .1 ~ A 123 z.» 0; ‘ l. ‘o 3% name-0 130 1. g; '0 P- 2'- ‘ V 140 .94 -1 '0’ 1n 2 1.00 0 9 103 300 f) 2 - .40 0 4 170 .100 "‘ 3 > ‘ o o u 5 9 10’ ° . 3 9 C 4 a — 10 z y 'I.“ S o o B u U 10' )— -1 10’ 8 a: u o a 0...“) a :00 non .0 1 I l 1 l L l 1 .02 1o" 10" 1 10 10’ 10’ 10‘ 10‘ neoucco FREQUENCY, 1.1,- n,u (RAD/S) Figure 2.1.--Reduced dynamic viscosity and elastic modulus vs. reduced frequency for SBS 7-43-7. Reference tem- perature is 138°C. Data of Gouinlock and Porter (1977). 27 associated with long-range motion of the chains of these block copoly- mers is not characterized by the peak of the loss modulus G"(w) but at a lower critical frequency where domain phase behavior dominates. Perhaps the most detailed account of rheological study per- formed on block copolymers is the IUPAC commission study of SBS melt compiled by Ghijsels and Raadsen. Steady, dynamic, creep and elonga- tional flows were conducted. The SBS specimen under study consisted of cylindrical polystyrene domains (18% wt) dispersed in the poly— butadiene matrix. The effects of pressure, temperature, and time between measurements on material properties were also tested. Their results can be summarized as follows. 1. The melt viscosity of the triblock copolymer is much higher than that of otherwise similar random copolymers of same composition and molecular weight. 2. The viscosity at low shear is very sensitive to shear history. 3. In the low shear region, the complex viscosity is as much as three times higher than the steady-shear viscosity at equal values of frequency and shear rate. 4. A residual shear stress depending on previous shear conditions is observed in shear stress relaxation experiments. Similar flow behavior, especially at low shear rates has been reported by Cogswell and Hansen (1975) with ethylene polypropylene 28 copolymer melt and Mundstedt (1981) with ABS graft copolymer melt. 2.2 Optical Studies Electron microscopy and x-ray difraction have become invalu- able tools in structure elucidation of block copolymer systems. Since the method of sample preparation is known to affect rheological results, a brief review will be outlined on how various workers have utilized the above techniques to identify factors affecting the mor- phology property relationships of block copolymers. Pedemonte et al. (1975a and b) have performed a detailed study of the dependence of their morphology and stress properties on the preparation of samples. For Kraton 1101 (SBS with 33%S), they have compared the original copolymer with films cast from toluene solution at two different evaporation rates (ca. 20 and 0.5 cm3/h), compression moulded films, and extruded and extruded-annealed speci— mens. From annealing studies, it has been concluded that the original material contains rod-like polystyrene domains. From the comparison of the electron micrographs and stress-strain curves of both extruded and extruded-annealed samples, the following conclusions have been drawn. The high values of the Young modulus are caused by a high degree of orientation of the polystyrene rods along the extrusion axis; the yield point is explained by the presence of many disloca- tions and thin ties which link consecutive cylinders. In the case of solution cast films, the morphology of samples prepared at a high evaporation rate does not show any regular arrangement of the 29 polystyrene which seem to have a rod-like shape, while for low rates, a morphology similar to that of'Uwaoriginal annealed samples is observed. In moulded films, the polystyrene chains form rod—like domains in a rubber matrix, but no particular orientation of the cylinders exists. But Lewis and Price used X-ray diffraction and electron microscopy to compare two Kraton 1101 samples--one, prepared by compression-moulding and another, a film cast from dilute benzene solution. They observed an anisotropy of mechanical properties with the former samples and an isotropy for the latter samples. Kawai et al. (1968, 1969) have studied films of SI copolymer of different composition obtained by evaporation of about 5% toluene solution. Electron micrographs of sections perpendicular to the film surface have revealed five types of morphology: (1) spheres of PI randomly distributed in a PS matrix for a PS content of 73 wt%; (2) cylinders of PI randomly distributed in a PS matrix for a PS content of 65%; (3) a rather disordered lamellar structure for PS content of 49% and 43%; (4) cylinders of PS randomly distributed in a PI matrix for a ps content of 33%; and (5) spheres of PS randomly distributed in a PI matrix for the PS content of 18%. The authors have also studied the effect of the nature of the solvent using one good solvent for polystyrene (MEK) and four good solvents of poly- isoprene (cyclohexane, CCl3, n-hexane and iso-octane). With the copolymer in such solvents, electron microscopy has revealed dis- ordered structures. These results contradict those obtained by slow evaporation of the solvent (MEK, dimethyl ketone and toluene) from 30 mesophase of SI and SIS copolymers (Gallot et al., 1969) of both lamellar and cylindrical type. A possible explanation of the dis- ordered structure observed by Kawai would be a too high evaporation rate fixing the disordered structure in the dilute solution. Kawai et al. (1968) have also tried to relate the composition of SIS copolymers to their morphology and mechanical properties. Polystyrene spheres were found dispersed in a polyisoprene matrix for a polystyrene content of 9.5%, slightly curved PS rods arranged nearly parallel in the PI matrix for an S content of 23%, a rather disordered lamellar structure for a PS content of 47%, PI domains of various shapes and orientations in a PS matrix for a PS content of 72%. Kawai et al. have also observed a systematic change in the stress-strain behavior with the copolymer composition, a change rang- ing from the behavior of a soft rubber vulcanizate to that of a carbon-filled rubber vulcanizate and finally to that of a hard, but toughened, plastic exhibiting a well defined yield phenomenon when the PI content of the copolymer increases. To explain the existence of three types of domain structures (spherical, rodlike, and lamellar) in SI, SIS, and 131 block copoly- mers cast from dilute solution, Kawai et al. (1969, 1977) have assumed the formation of micellar structures at a critical concentration during solvent casting. They have proposed an analysis of forma- tion of three types of domain structure and the size of the domains taking into account thermodynamic and molecular parameters such as incompatibility between the PS and PI blocks, total chain length and 31 weight fraction composition of the copolymer, solvation of the blocks and temperature. They conclude that the block segments are preferen- tially oriented along the direction perpendicular to the interface between the two phases and they postulate that the micelles formed at a rather low concentration maintain their structure in the solid state without reorganization. During evaporation, the micelles shrink in the direction perpendicular to the interface between the domains. Spherical micelles shrink isotropically while rodlike and lamellae micelles shrink anisotropically. In rheological eXperiments increasing attention to sample preparation and morphological char- acterization are being given as attested by the works of Kraus and Rollman, Gouinlock and Porter and Ghijsels and Raadsen. All these studies tend to illustrate the basic feature of block copolymers, i.e., the additional complication that arises from the constraints that restrict the components to separate regions in space. A more complex picture is further introduced by the geometry of these domains which may contribute to anisotropic deformation. We avoid the latter difficulty by choosing a block copolymer with spherical domains and treat them as elastic barriers. CHAPTER III A TRANSIENT NETWORK MODEL FOR POLYMERIC MATERIALS WITH SPHERICAL MICRODOMAINS 3.1 Objectives 0n the basis of rheological experimental observations presented in the previous chapter, we undertook to formulate and test a kinetic network model based on network theory for block copolymer melts with spherical microdomains, incorporating realistic and tractable rate terms for attachment, and detachment of segments (flexible sub—chain) and domains. The resulting segment distribution is non-Gaussian so that a general expression proposed by Yamamoto is required to calcu- late the macroscopic stress. In the following chapters, this model will be tested fortufiaxial extensional, simple shear, and oscillatory flows in both steady and unsteady conditions. Next rheometric data shall be presented on a well characterized diblock copolymer sample --poly(styrene-b-butadiene) whose morphological structure is known and the material functions will be compared with model results. 3.2 The Rate Terms Figure 3.1 depicts spherical, rubbery domains uniformly dis— tributed in a soft, continuous phase. The position p-is referred to a fixed origin while the position_r is referred to the end of a seg- ment which may or may not be at a domain; R_denotes the nondimensional 32 33 1:0 :1: '0— Figure 3.1.--A polymer network with rubbery domains. 34 position r/Nl where Nl is the extended length of the segment with N subunits. An active network segment in this representation is a flexible strand bridging rubbery domains and/or entanglement junctions in the soft phase. A segment distribution function f(R,N,t) may be defined such that fdiR is the number of elastic segments in the network with an end to end vector in the range R toIR + dB at time t and composed of N subunits. This function obeys the evolution equation of Yamamoto 31;— + v - (_R_f) = 6(3.N) - B(B,N)f (3.1) where G(R,N) and B(R, N) denote the rate of creation and the coeffi- cient of destruction of segments with N subunits; R_denotes the velocity of such segments which may be expressed following Phan-Thien and Tanner as B=( Ill"— - a9) '5 (3.2) where E is the velocity gradient and Q the deformation rate tensor in the fluid; E is a slip coefficient. A flexible segment in this representation may be constrained by impenetrable barriers at one or both of its ends, as in a diblock or triblock copolymer melt. Hesselink (1971), Napper et al. (1975), and Edwards and Dolan (1975) have derived one dimensional equilibrium dis- tribution functions for such segments, taking the presence of these barriers into account by imposing the boundary condition 35 r = 0,11) = o (3.3) at the domain boundary. With spherical domains, the spherically symmetric form satisfying Equation (3.3) proposed by Chompff may be used. T-- a2 —- 3 r2 fe (r,N) m exp exp - §-——— (3-4) ~ rZ/Nt2 N22 This distribution is originally attributed to Reiss (1967) and Yamakawa (1968) who proposed a general expression for the total potential energy, E, of the configuration of a free polymer chain as N 1 E = 2 u v=1 1 . . +-—Z V (r..) (3.5) 1,1+1 21¢j 13 Here the monomers constituting the chain are treated as hard spheres distinguishable by their positions in the sequence constituting the polymer and are held 1n place by ass1gned potentials ui,i+1(ri,i+1)‘ The spherically symmetric interaction potential between monomer i and monomer j is represented by V(rij) where rij is the distance between the centers of monomers i and j and N represents the number of monomers in the polymer. The configurational partition function for a polymer molecule whose first monomer (segment) is fixed with its center at the origin assumes the form Z = J . . . J exp (- E/kT)dtzdt3 . . . dtN (3.6) 36 where (he, dT3, etc. are the volume elements for the second and third segments, etc. The integrals extend over all space. The configura- tional partition function for a polymer molecule whose Nth segment has its center fixed a distance r away from the fixed first segment assumes the form Z(R) = J . . . J exp (~E/kT) die, dT3 . . . dTN_1 (3.7) where in the integration it is understood that the first and last segments are a distance r apart. The authors then calculated for the configuration probability of a free chain in which one end is fixed and is constrained so as to decouple the many body problems. This can be represented as 2 - <1 P ~ eXP ( 23:2) 9(1) (Jig-l) (3.8) Here the function ¢N(r) represents a spherically symmetric external field (centered on the first segment to which the Nth segment is subject and clearly depends on rN' The spherically symmetric form of ¢N(r) of Equation (3.4) adopted by Chompff predicted very well the stress-strain relationship of rubber vulcanizates at high extensions. For block copolymer systems, the parameter "a" in Equation (3.4) describes the range of repulsion between continuous elastic segments and the domain to which they are attached. If the number of segments attached to a domain is small so that the range of repulsion between segments is less than the maximum end-to-end distance of the segment, 37 a < 1. The creation rate expression G chosen in this study is pat- terned on Equation (3.4) and written as 23 3/2 2 2 e (3,11) = c Egg—(3%) exp[- i}; - 3N2R] (3.5) where C is a constant rate coefficient. The symbols r and R denote magnitudes of the vectors 3 and 3 respectively. The shape of the distribution in Equation (3.5) is shown for several values of a in Figure 3.2. With increasing a, the peak shifts to higher values of R, i.e., the end-to-end distance of most probable segment is increased. From the previous discussions, the repulsion coefficient "a" is an inverse function of temperature. The applicable region of temperature for G is T < T f-Tt where T and Tt are the block 9 . 9 copolymer glass transition temperature and transition temperature to a single phase respectively. Here "a" has a range of 1 < a_: 0. A consistent expression for the rate coefficient of destruc- tion 8 is obtained from the relation 8 (R,N) = Bo[1+€(A(R,N) - A(O’N))/kT] (3 6) N where the leading term is the contribution from Browian motion and the second term is associated with the change in entropic free energy A of a segment in the network by flow and repulsive interaction. (Acierno et al., 1976). Writing the configurational partition func- tion Z in accord with Equation (3.4) as r' Z (R,N) = K0 exp L:-% NR2-a2/NR{] (3.7) 38 .mcowuoczm :oepznwcpmvu cmwmmzmw1coc cme_mELozu1.m.m wcsmwa a Ho. 80. 93an o [gun/,9 - z/zaus-Idxa 39 and using A = -kT an (3.8) where kT is the Boltzmann's temperature, we obtain 2 2 B(R.N) = 3.0 +13%— + EL) (3.9) NR2 The rate of destruction is Bf so that at R = 0, Bf = 0, since f is an exponential function of 1/R2 while B is a polynonnal of l/RZ. Thus the rate expression is well behaved. The initial distribution of segments is given by f(B.N,t=0) = G(R,N)/B(BJN) (3.10) The moment integrals are considerably simplified if it is assumed following Fuller and Leal. that 8 << 1 so that f(B,N,t=0) 3 G(R,N)/Bo (3.10a) with e << 1 and a < I, the third term in equation (3.9) is clearly much smaller than the other terms. The rate expressions outlined here should be appropriate for a block copolymer or a filled polymer melt containing spherical domains or particles with low surface denSity of segments and high interpene- tration in the continuous polymer phase. The elastic free energy of the network is largely in the flexible segments of the continuous phase. 40 3.3 The Macroscopic Stress Tensor As already mentioned in the preceding section, the rate expressions chosen here will lead to a non-Gaussian segment distribu- tion f; so a general equation of Yamamoto is used to find the macro- scopic stress S in the network _ 1 91: (BM 2 - 'p dR .BBf(B)N9t)dR (3-11) OY‘ _ 1 51/1 (KNEE, é - < R dR , (3.11a) Combining (3.7), (3.8), and (3.11a) yields § = 3NKT <.BB.- 2a.BB. > (3.12) 3N2R“ The validity of this model is examined in the following two chapters with detailed stress calculations for uniaxial extensional, simple steady shear and oscillatory shear flows. CHAPTER IV PREDICTED STRESS BEHAVIOR IN EXTENSIONAL FLOWS 4.1 Uniaxial Steady Extensional Flow The kinematics of this flow are described by V1=TxaV2=-%y:V3=-%Z (4.1) where F is the magnitude of the strain rate. The steady deformational rate tensor ;* is given by g=r 0 ~11 0' (4.2) 0 0 -._J where T is the magnitude of the effective strain rate experienced by the network I = T (I-E). Since L* is a diagonal tensor independent of R, segment evolution equation of (3.1) becomes A .81“: .211. at ia: 2 3y 2 32 +Fx _ __i. at 8x 2 G(R,N) - B(E,N)f (4.3) This hyperbolic first order partial differential equation has three characteristic lines as shown: x = x0 exp (It) _ i y - yo eXD (- 2t) (4 4) _ I: z - 20 exp (- 2 t) 41 42 Using the macroscopic equation of stress, (3.12), the primary normal stress difference N1 in uniaxial extension then turns out to be : _ _ 2 - 2&2 2 2a2 . N1 ‘ Sxx Syy ' 3N“ L< X (13N2R‘*)>’ < y (1' SAP—Rip] (4'5) The two moment integrals in (4.5) may be evaluated using transforma- tions described in Appendix A, similar to those employed by Fuller and Leal. Defining non-dimensional time, strain rate and elapsed time T = Bot; T = T/Bo t'==BO(t-t') (4.6) we may write n kT T 1 “1(T’ = 132a I [11(1) - 12(111e'T + ) e‘T [11(1') - 12(1')]d1'} (4.7) where no E C/Bo; and 11, 12 are integrals over space in spherical polar coordinates as noted in Appendix A. The integration over one of the angular coordinates, w is carried out numerically, avoiding a singu- larity at p = n/Z with a generalized Gauss-Legendre quadrature formula of Krylov (1962), for 12. 4.2 Results 4.2.1 Steady State Stress At steady state equation (4.7) reduces to nokT . N1 = TIE—a— emT (Il-Iz)dT' (4.8) 43 Both 11 and 12 depend on the two parameters a and e. If both a and e are set to zero, Lodge's rubber like liquid model is recovered. With a alone set to zero, equation (4.8) may be written with a damping function h(I,T)and a strain measure B(T,t) in the form proposed by Wagner (1979a) N1 = noKT [de' e-Tl h(T,t) B(i,t') (a=0) (4.9) 0 with h(I,T'_)3 (141—;a (1-e'."T'))‘2 (1+ f1: (e2fT|-1))-3/2 (4.10) and B(I,t')E erT' - e-fT| +-§: (2e2le + e'fT' - 3) (4.11) Such a factorization is not possible for the case where both a and e are nonzero, and the distribution is non-Gaussian. The normal stress difference N1 may be scaled with nokT--a shear modulus--t0 compute a dimensionless elongational viscosity at steady state * Nl/nokT N1/? = (4.12) fit: I 00 Figure (4.1) presents a comparison of elongational viscosity plots againststrain rate calculated with a fixed value of e = .01 and several 44 OOH .Ho.o u 0 saw: =m= gmpmsmgma covmngmg Co #0040“ .mpmc sweepm mmm_:owmcmswu .m> zp_moomw> chowmcmpxm umNTFmsL0211.H.¢ m;:m_a .._ OH H H. Ho. _ . q H 0.0 m0. m." M 1 1 OH we . _ . OS 45 values of “a." With a = 0, the elongational viscosity levels off around a dimensionless strain rate of 0.1 to a value of 3--the Trouton ratio between the low strain rate values of extensional and shear vis- cosities. As the value of a is increased, an upturn in viscosity is noted in the lower range of strain rates; an apparent yield stress may be identified at the lower strain rates on each of the plots with a f 0. It is worthwhile to point out here that in the limit of zero strain rate, N1 is zero and the elongational viscosity is finite; this must be true of kinetic network models such as the one discussed in this work. An analytical expression may be obtained for the apparent yield stress at low strain rates by simplifying equation (4.8) for i << 1. Ny 4 a 5 ° ' nokT = §1+2a (1+ 2 8) (F << 1,1” 7' 0) (4.13) It is readily seen from equation (4.13) that with 6 << I, the apparent yield stress is much more sensitive to the parameter a. Recalling that the value of a is directly related to the range of expulsion between segments attached to a domain, this relationship between the apparent yield stress and the parameter a is reasonable. The signifi- cance of this parameter is further illustrated with the elongational viscosity data reported by Munstedt (1981) on ABS block copolymers at 190°C with various concentrations of butadiene, the rubbery component. The apparent yield stress NY from the data is tabulated against rubber concentration in Table 4.1 along with the non-dimensional yield stress 46 Table 4.1.--Model parameter ”a" from data of Munstedt % Butadiene Observed Yield Stress N /G*- Estimated in ABS NY (Pa) Y 0 a 20 2.0 x 103 .004 .005 30 5.0 x 103 .010 .013 43 1.5 x 104 .030 .038 *G0 Plateau Storage Modulus of 0% Butadiene in ABS. and the corresponding values of the parameter storage modulus obtained 5 Pa. This table shows from Figure 20 of Munstedt's paper as 5 x 10 that increasing rubber concentration in the copolymer is described by increasing values of a in the present model, so that the segment distribution is increasingly non-Gaussian with higher concentrations of the rubbery domains. The effect of the other parameter e is more noticeable in the peak elongational viscosity attained at dimensionless strain rates of order 1. This peak is lowered and moved to lower strain rates with increasing values of e, as shown in Figure 4.2, where plots of elonga- tional viscosity are presented with a fixed at 0.05, but with several values of c. This trend is understandable since 6 is a measure of the dependencecfi function destruction cu) the deformation. Data are not available on peak elongational viscosities for block copolymers to verify this trend or allow a quantitative comparison. The effect of a on the peak value is only slight; increasing a leads to a small reduc- tion in this value as seen in Figure 4.1. 47 ooH OH .mo.o u m ;p_z =0: gmmemcma cowpozgumm m *o pummem .mpmg cvmcpm mmm_co_mcmswu .m> auwmoomw> PacowmmmwMW11.m.e wcszd OL-o Hc. moo.u mo. Ho. ooH 48 4.2.2 Stress Transients The development of stress in experiments with a sudden step in elongational strain rate, T may be predicted with the help of equation (407') at several values of F.. The results are plotted in a ratio N1(T)/nokTT against T in Figures 4.3—4.5. In Figure 4.3 a is set to zero and at I = 1 and i = 10, increasing 8 leads to reduced overshoot. Figure 4.4 presents the transient elongational viscosity at F = 0.1 with e = 0.01 and several values of a. As a is increased, the trnasient viscosity is increased at all times. At T = 1, however, as shown in Figure 4.5, the transient elongational viscosity curve changes only slightly as a is increased. The data of Lobe and White (1979) on carbon black filled polystyrene melts at 170°C (see Figures 5-7 of their paper) show similar trends with concentration of filler at low elongation rates of 0.0063 sec-1 and 0.02 sec-1, increasing carbon black content leads to higher transient elongational viscosity at all times. Once again, the value of a in the present model corre- lates directly with the concentration of filler in the material. 49 100 I c =.005 E =.OOS '01 .01 10 - ”111+ .1 1 .1 1 8 T Figure 4.3.--Normalized transient extensional viscosity vs. dimensionless time as a function of strain rate. Effect of the destruction parameter c with a = 0. 50 Figure 4.4.--Normalized transient viscosity vs. dimensionless time as a function of strain rate. Effect of the repul- sion parameter "a," with F = 0.l and e = 0.01. 51 100 i *+ ”E a-.3 0 10 - ‘ 1 F ‘ l 1 7 Figure 4.5.--Normalized transient viscosity vs. dimension— less time: Effect of the repulsion parameter "a" with F = 1 and e = 0.1. CHAPTER V PREDICTED STRESS BEHAVIOR IN SIMPLE SHEAR FLOWS 5.1 Simple Steady Shear In uniaxial steady shear flows, the effective deformation rate tensor is given by 0 2-5 g = 324 «a 0 0 (51) _0 0 0) where 1 is the magnitude of steady shear rate. For convenience, this 1 tensor is diagonalized‘ by introducing a tensor T such that T- L*T = V Where V is a diagonal and r— . é. . _% .— -1(2-E) 1(2-5) O .1 : /7TT:§T O O O L _ Next a coordinate transformation leads to a new frame 5 = p(p,n,z) such that R = T - p. The diagonalized tensor V is composed of the ~ eigen values of the tensor L* and expressed as 52 53 r1 0 0— v = 12"1 0 -1 0 (5.3) 0 0 OJ where ~ ° 4 . m = Y[g(2'€)] s 1 : H0 The evolution equation for segment distribution becomes .81 ~ at _ ~ .31 = " _ A 3t+'mqg- min—8.71 GWJU BQJHf (54) where 6(9):) = (sq-9,11) (s 5) B (p,N) = B(I-Q,N) (5 6) In terms of the transformed coordinates T “=0 I'I'B (57) R = x + y + z = p2 - 2an + n2 + 22 54 The characteristic lines for Equation 5.4 in terms of p(p,n,z) coordinates are = poe1mt/2 -imt/2 n hoe z = 20 (5.8) Applying the macroscopic equation of stress given in (3.12) for a non—Gaussian segment distribution function f, tangential and first normal stress difference relations can be generated in terms of moments in cartesian coordinates as S = 3NKT < (1 - Z§E__ ) x > (5 9) xy 3N2R4 y ' I _ = 2 _ 2 _ 22 2 4 N1 _ Sxx Syy 3NkT [<(x y ) (1 Za/3N R )>] (5.10) Necessary cartesian components of the stress tensor can be evaluated from the transformed coordinates f(p, n, 2) using the expression xx = 9 . IT(1 . 9) (5.11) The cartesian moments are related to the moments of the transformed frame by the multiplicative factor, det(T) and these are expressed as: fil__§ ‘E 2_. 2 -§§ 2 (p n > (5.12) (1-5) ='——-——;§ <0 - 2(1-€)on + 02 > (5-13) Moment integrals in transformed coordinates are solved in Appendix B, first by evaluating Equation (5.4) through the use of transformations prescribed by Fuller and Leal. 5.2 Results Results of steady state dimensionless viscosity (gxy/i) and first dimensionless normal stress difference, N1 obtainable from Equations (5.9) and (5.10) can best be discussed with and without "8" equal to zero. With "a" equal to zero, a case where the initial distribu- tion of segments is Gaussian the steady shear viscosity and the first normal stress difference are obtainable from Equations (5.19) and (5.10) as: oo s _. _ -T' dT'(Slan' - E(cosmt' - 1)) 5(1) = J 2 2 e m 2 1 3 2 (5.13) - _ + Y (1 g) 0 (1+2—rE-fis'inm'f' - 2%— (COSTllT'-1) ' (1%)2) / m ~ A 2 GD_T. dT'(-§(2-€))%(1+ET' - cosmt' - e/m Sin mt') ”1(i) = 1-g e 28 252 1+5t' 2 3/2 (5'14) (1+_fi sin mt' - 2 (cosmt'-1) - (szf—) ) o m Here n and N1 are defined as 56 - 5x /1 n5 NOkT/B0 s N1 N : 1 ‘ nokT where 5 E y/eo is the dimensionless shear rate, n0 is the initial concentration of network segments and 1' denotes the dimensionless elasped time, 80(t-t'). Further, if e is set to zero in Equations (5.14 and 5.15), viscometric material functions similar to those of the Phan-Thien and Tanner model, [see Equations (30) and (31) of Phan-Thien and Tanner, 1978] result fi = (1'5) ~7~ (5.16) 1 + €(2-€)Y NI = 311‘5)i 5,_. (5.17) 1 + 5(2-€)i The non-linear dependence of shear viscosity on shear rate in most polymeric systems is accounted for in this model through the slip mechanism, E. In Figure 5.1, the effect of the destruction coeffi- cient c on dimensionless viscosity is presented as a result of com- puting (5.13) using a 40-point Simpson's composite formula. This result shows that "5" does not affect the trends in viscosity vs. shear rate, but merely changes the scaling factor, nOkT/BO. Similar 57 0.3 .o u m =0: .memsmgmq cowpochmmn mgu 4o pommmm .mp8; cmmgm mmmFCOVmcmec 4o covpoczm a ma xpwmoomw> cmmzm vawgmscozu1.H.m wezmwa To 04 CC) I! w conclusions have been arrived at by Phan-Thien and Tanner as well as Fuller and Leal. Since the value of e affects only the scaling factor, subsequent curves of viscosity are plotted only for e < 0.01 so that nokT/B0 coincides with the zero shear rate viscosity. For the non-Gaussian model (a f 0), steady shear viscosity and first normal stress difference expressions are derived from Equations (5.9) and (5.10) as ~(i) = (eza ))(-g(2- i e-T'(-h(: ') B (7 ") I) Y 1+Za L (1-€)2'? YsT 10 YsT 0 66:62 7 T +'*_—_2hl(Y’T') 811(y,t'))dt'} (5.18) m(l-E) where 118,10 . (82-5)): 3/2 (1 + 2 E-s'n ' - 282 (cosm '-l) - (1+ETRZ) ' m l 1111' -—m—2' T 1‘15 ) 810(i,t') = sinmt' --% (cosmt'-1) 811(i,t') = 1 -E(2-E) cosmt' - (1—E)2coszmt' - mt'sinmt' ~ 1 l 2 1~l (1+2fi'5ian' - 2E(cosmt'-1) - (ifETY32C(y,T') 3») 41(1) (£3 )Tiit) dt'e- (hm ) 3208,. ) + 88 a2 O x (1-5)h1(i,11821(i.r')) (5.19) where B20(i,1') = 1 - cosmt' --% (sinmt' + mr') B21(i,t') = E:f:?§ -t' - sinmt'( (I-E)3 + (COSTEé-ll - 1) The second term in (5.19) is an additional contribution to the stress level of newly formed chains with respect to the degree of their repulsion from the domains. Again using a 40-point Simpson's composite formula Equations (5.18) and (5.19) are computed. In Figures 5.2 and 5.3, the viscosity and first normal stress difference are presented with e = 0.005 and various values of "a." For a i 0 Figure 5.2 shows an apparent yield stress and a quick decay of viscosity to a "plateau" at dimensionless shear rates of order .01. At large shear rates the model then yields the power law behavior. The general shape of the curves in Figure 5.2 60 .mo.o u w .moo. n o =.m= Lmumsmgma cowmpzqmg 4o pomw$m .mcowpocsm mocmgmmw_u mmwgpm Peace: “mew; vawmeLoc use cowuoczw zpwmoumw> cmwgm cmNPPmEL0211.N.m mesm_a S o; I. 5 S. no _ _ . _ s _ _ to 1 i ... OJ r1 \ luoé .. \\ 1!- . L 14 J 1 o 21 0.0 1 Ines m5 1 m8 1 a 2C 61 .moo.o u a .mo.o n a :3: 250580 aw_m 5:0 to Soweto .mcowpoczw wocmgmmewu mmmcpm Peace: pmcwe use xawmoumw> cmmzm.umepmeLoZ11.m.m mczmva o.oH o.H w Ho.o Ho.o H.o . .Hd q a _ q _:qu3 .1-.. __.___J N . _.._nu N. m.o 1 H.. 1 n t mo.o u a n: 1 § 11 o.H 1| L'snnnul1 14 12 L N.o 1 o.oH11 H.o 11 . modnw .. .4 H.o o.H o.oH 62 are in good agreement with flow curves of block copolymer melts and even with those of triblock melts. There is sharper upturn of shear viscosity at low 5 with increasing "a” and experimentally a sharp upturn of shear viscosity is also noted at higher fractions of the domain phase. Thus "a" correlates directly with the concentration of the domain phase. The upturn in shear viscosity results is not as drastic as those shown in the extensional flows. This is FT for attributed to functions controlling their strain measure e extensional flows and sin mi' for shear flows. The effect of the slip factor as shown in Figure 5.3 is to change the power law behavior of the material that occurs at large deformation rates. The model does not predict any new trend in first normal stress differ- ence except a slight increase in magnitude at all shear rates, as compared with the Gaussian model. The large difference with high slip ratios at large shear rates (see Figure 5.3) are predicted even with a = O. The trend in normal stress-shear rate relationship pre- dicted by the model awaits further evaluation by experimental data. However, literature is devoid of such data for block copolymers mainly due to general difficulty in collecting reliable normal stress data in conventional rheometers. The normal stresses of all melts of high M.W. is difficult to measure due to the compliance of the instrument at high shear rates. Such problems encountered also in this study will be discussed in the experimental section. From the results of extensional and shear flows, the contribution of the non-Gaussian nature of chains occurs at small chain extensions. It 63 10.0 1 1441111] I J 1141J Figure 5.4.—-Normalized transient viscosity function. Effect of the repulsion parameter a, e = 0.005, g = 0.05. 64 cowm_:gmc mg“ mo powwow OH .mo.o u w .moo.o u u .m u + =.m= gmmemcma .cowuoczw xpwmoomw> pcm?mcwcp umNPPMEcoZ11.m.m mcsmwd 0.0 m.o H.o o.H 65 _ 11k ~ Y = 10 N1 "’ 4—— =3 1.01—- ,- i- F l. OJ. 4 lJ-llllLl 4 JLJlnLi 0'1 1.0 T 10 Figure 5.6.--Normalized transient first normal stress difference function a = 0.0, e = 0.005, E = 0.05. 66 is worthwhile to emphasize here that the ensuing newtwork model is mainly applicable to the low deformation region. Transient stresses are computed by using time dependent moments as developed in Equation (8.9) in the stress expressions of equations (5.9) and (5.10). The results are plotted as fi+ (T) and N1+ (T) vs. dimensionless time T in Figure 5.4 to 5.6. Figure 5.4 shows the shear growth viscosity at low shear rate as a function of time. The magnitude of this material function increases strongly as the parameter “a" increases, growing monotonically with time until it reaches the steady state value. In contrast to the IUPAC data on SBS melts, no stress overshoot is predicted by the model until shear rates of the order 1 as shown in Figure 5.6. In these data 1 stress overshoot was noticed at shear rates as low as 0.01 s' . The strain at stress peak, it has average value of 3 at § ~ 0(1) and max increases linearily to 6 for instance at v = 30. As g approaches 0.2, §tmax stays fairly constant at 3. Many workers, Osaki et al. (1967), Graessley et al. (1977) have correlated this data to the total strain on the material. An experimental value of thax = 3 have been reported by the former researchers on homopolymeric melts. Regardless of the value of a, the stress growth curve predicted is oscillatory when the slip ratio 5 > 0.1. More recent experiments (Osaki et al.) have discountenanced the presence of the undershoot after an initial overshoot. 67 5.3 Oscillatory Shear Flow In oscillatory shearing, L* is given by F0 2-g 0‘1 Y E* = EQ'COSwt -g 0 0 (5.20) L o 0 OJ vmerei = wYo and w is the frequency of oscillation and yois the strain amplitude. Exactly the same coordinate transformation used for steady shear flows is applied here to obtain the specific evolution equation as 3f 170 3f 1m ST A A -5— +-—§— (coswt)(p 56) - —§9-(coswt) (n 55) = G(91N) - B(BaN) (5~21) where m0 = wYO/El2-g) We next make a domain transformation in the independent variable t to u such that u = sinwt to obtain the characteristic lines as shown in (5.21) . This simplifies the computation of f as described in Appendix C. p = p eimOu/w 0 4m u/w - o 0 oe z = z (5.22) 68 An oscillatory shear stress can be obtained through the macroscopic stress equation of (3.12) in terms of moments in the cartesian coordinate as S = 3NKT < (1 --—g§E——) x > (5 23) xy 3N2R4 y ' We note the velocity gradient is varying sinusoidally with time; hence, the shear stress varies sinusoidally after transients have died down and may be represented as _ Toot SXy — Re (Soe } (5.24) where S0 is in general a complex functionoFOr small strain amplitudes a strain independent complex viscosity n*(w) may be defined as the limit Tim In* (w,y )I= lim S /wv =|n*(w)! (5'25) +0 0 +0 0 0 YO YO where n*(w) = n'(w) - ln"(w) The real part of the complex viscosity, n' (the dynamic viscosity) is associated with energy dissipation and the imaginary part n", wn" = G' is associated with energy storage; these are the so-called linear Viscoelastic moduli and are related to the oscillatory shear stress by 69 Sxy = n ($05wt + n v0 Slnwt (5.26) Upon computing the moment integrals encountered in equation (5.22) as shown in Appendix C, we obtain the following expressions. ~ 26. _ l n'(G,YO) = §+2a (:-:) dT'e T [EO(®,T')BO(0,T') o + 6a2h1(&,vo,t') 81(@.I'{] (5.27) n“(' ) = 92a LZ:§l_. dt'e'T' h (m 1') B (0 t') + “’Yo 1+2a (1_€)2 __o ’ 2 ’ O 6a2h1(&3,v0,1') B3(J1,T')] (5.28) where 0 = 00/80 n =BOn /nOKT n =Bon /n0kT 1-(Hauflotfi 2 2 . ~ . 2 1 _ (1+ET'22 + 462 (0T'(1-%?3 +‘%?'S1an ) 2 2~ (1-g) mowz C (J) ,YO’TI): 70 BO(&,T') = 1-cosfit' +-§—(&T' - sin01') w _1_ h1(E),Y :1") = (‘512-§))2 2 0 C(0,T')%(l + mT'(Z-m0) + mg sin 01') Bl (6,1') = (Qt' COS&T' - Sin&T')(1+ gELL—2) (l-E) BZ(&,T') = SinQT' +-% (1 - COS&T') so 0.) A 82 v 1—1 23 II (1-1111'51'115131' - cosmT')(1+ Ell—2) (r-E) Equations (5.26 and 5.27) involve m0 = vo/§(2:E) in their second term making fi' and fi" dependent on the strain amplitude. However, numerical analysis of these function at v0 < 0.1 showed no significant difference from the linear results. In this region then, it is assumed the linear response applies and thus compute the complex viscosity function as |fi*(0.io)|§lfi*(0)l = (fi'2 + a"_m>wuomqmmc mum; cmmgm can xucmscmgw we mcowpucsw mm prmmum_> acmmpm vcm owsmcxu meanu ume—msgo: mgh11.n.m weaned w .3 .3 .2 T2 1 N.2 o u u - Hoo 4 o.H [Hollllllll rullrlll 1llll mo.o n w moo.o n m m.o n m .C IIIIIIII TE ||I1 - n - o.oH L3 ‘19 ‘l‘l‘gl 73 flow (i.e., y = 0). this can easily be computed with the explicit distribution function given as e‘8(psnazsN)t f(o.n.z,t) = f0(o,n,z) A _B(010,Z:N)t + 9(010121N) (1'9 ) (5.29) B(o,n,z.N) CHAPTER VI SAMPLE CHARACTERIZATION AND EXPERIMENTAL TECHNIQUES 6.1 Material and Sample Preparation The block copolymer employed in this investigation was a research grade poly(styrene-b-butadiene), C0326-9 (containing a small amount of an antioxidant, Ionox)generously provided to us by Dr. Lu Ho Tung of the Dow Chemical Company. Characterization information for this copolymer is provided in Tables 6.1 and 6.2. Approximately 0.2 cm thick copolymer films were prepared for both rheological and morphological studies by the solvent casting technique (Hashimoto et al., 1977). Thin films of the copolymer were made by dissolving 20 gms of copolymer in 100 ml of toluene and the solution transferred to 10 cm Petri dishes. These solutions were then placed in a vacuum oven kept at 30°C with all port outlets closed except one connected through a valve regulator to a hood chamber to insure slow evaporation. The oven was periodically flushed with nitrogen to prevent the oxidation of unsaturated bonds in the butadiene phase. After the films were visibly dry a procedure requir- ing five days, they were further vacuum dried at 80°C. It was assumed that adequate drying was achieved when the decrease in weight of the sample varied by no more than 0.005 gm. Again to prevent sample degradation during weighing, the vacuum oven temperature was 74 75 TABLE 6.1.--Block copolymer characterization Specimen T B 81°Ck 5 Block Code ype -fi -fi -fi Ht. Percent — -— -— N w/ N B Block “N Mw/MN C0326'9 (S'B)1 10,000 -1.1 5.9 232,000 ~I.7 TABLE 6.2.--Property of glassy continuous phase Molecular Weight SolubilitybParameter Glass 1 . . Structure Between Entanglement (Cal/cm3)2 nggzigion C ture °C Polystyrene 33,000 (8.1) .05C 1009 aValue derived from Newtonian Viscosity data of linear polymer (Berry and Fox, 1968). bHashimoto. et al., (1974). CSolubility parameter difference between PS and PB. dKraus and Rollman. Note: Polybutadiene Tg ~ -90°C Block copolymer - liquid above 100°C. 76 decreased to 25°C and the sample allowed to cool in vacuum. There- after, the oven was brought to atmospheric pressure with the nitrogen flush. This procedure was repeated until the constant weight was achieved. The sample was further annealed at 110°C for 24 hours. The film samples were then placed in a vacuum dessicator and a representa— tive sample was used for structure elucidation by electron micro- scopy. 6.2 Electron Microscopy The domain structure of the film specimen was investigated by transmission microscopy in a Philips 201 electron microscope operated by K. Baker of Pesticide Research Center, M.S.U. After embedding in a Spurr resin, the film was presectioned, stained, and fixed with Osmium tetroxide, 0504. The specimens placed on a support were allowed to stand forabout half an hour at room temperature over a 2% aqueous solution ofOsO4 stabilized with a Sorensen phosphate buffer, in a small, tightly closed glass vessel. The stained films were then cooled with liquid nitrogen to approximately -150°C and cut on a Sorvall Porter-Blum, MT-2 Ultramicrotome with a diamond knife. Ultra thin sections of about 800% thick were cut normal to the film surface by the Ultramicrotome. Figurest1.and 2 show some of the typical electron micrographs of the butadiene-styrene block copolymer at different magnifications. 6.3 Morphology The dark areas of Figure 6.1 are the polybutadiene phase selectivity stained by 0504 while the white portion is the polystyrene 77 fl ' a a. 1 .90... 95f I O c I .00. ..a .. 0 .r I' 0;. r J .‘ .AP'... 9.. a 1‘“. 0.l. ._ r...‘ . 05w; .90... o . 0.: o .. uoflfl‘ .a um...‘ . ... .a. r .. a - 0.. a . .a (a s 0 so. n . u‘a‘.afltavl.b .‘1. ma! .I .uc-e...a......oo. .. 00:” yo...-’ woo... o..t a..\+ . to... 1...". a Oofis ‘10.... d a 2. . o. . ’a'. M V. 0.. 1“. ,..- .0. n .. a ”.191. '0'... . 4‘... unto 9:0 . I; A o. I. D! i n i A c... n. . 0: v 1. a- 1o s 0 x: O .. .a.. on On 0.. '0.- c O F. Uw.0..’..‘ C .J O...- ‘ C v . a in a o It a 19.. v 0 .0‘05 01*. I it... I a". ”TO 0 a... .a.0.1 C .fiu. ..u s n n V 1 1 .0 I l O o +. O 60 .6 a o 3.1;". so.“ 1)t;¢..vug.fi$-flas r “hooves :I a u . a .r.. a v . . . . 'Ip..9.m ‘0 as 0"... u O.fl o v I055. ’Oé.‘.: U. . .«w u . 1 p .. Inc-95 Q o 1 {no}. u A vw. .1 .02... .. 0. '3 a ”6. . U.- i! .4. '5. In. 0 .p.‘ O O .. ‘&.~D.h .‘1 n . l. “‘0” ‘. 2 w ”0 v .‘O .. . I. ...ur.t‘ “h i . 0 if. 5.. a '5 .1. without . 0 ad... .. In 1.....mJ6...’ 1.0. .c' a: O o; o u "1 ”to-4. o a raw U . 6 5A..” . 1: i .05.! ‘C. . V C. I 9. fl ’ n,t cad P.- .r O . b.091... « ‘M C. .. ofi its, 0. U .‘0 ’O .P I . . O I .”.~. . . 1W.” .1 1 .9 O I. . O. p 4 G a. «QC . O ‘3" \ i 4 to \o O 9 . C Q .0 ‘0“. ‘ n. #on o we no. it... .0 be b. 9.00 '0! a. . . .2. u . . .. n. .h 0‘3: at” N v 9QflrOIdsa. ll 1 ‘- o - H _ . .. - —‘ ‘.u..'_:.- 0. ~. ‘. .5....50“!.I\'v~...' ....i. .. N «- h I“ .0 o .. . . ’ $5£o.fl .0... 4.6.6.. .nv: 05.. .U’DQOaL‘ .O‘.‘ ~‘o.’ (has. on a} o . v.9 . on» be}... . to. 91.05%“. ‘PJafiKah. 01.1. ' fitmnpxod r 0...? “.4 to tat-.091- paw-11......powtb 1\.l.fibun_ t x 50,000. specimen a Figure 6.1.--Typical EM micrograph of ultra-thin section of poly(styrene-b-butadiene) 78 Figure 6.2.--Typical EM micrograph of ultra-thin sections of poly(styrene-b-butadiene) specimen at x 150,000. 79 phase. The absence of spherical bundles or lamellae structure indi- cates only spherical microdomain structure of polybutadiene uniformly dispersed in a matrix of polystyrene blocks present in the copolymer specimen. The spherical domains have an average diameter of 350A and an average interdomain distance of 500A. The thickness of the domain boundary interphase, AR directly related to the degree of compatibility of the blocks is indeterminable by electron microscopy, but are known to be significant for low to moderate M.W. copolymers such as this specimen (Leary and Williams, 1970; Krauss and Rollman, 1976). Hashimoto et al. using SAXS studies have reported AR values for S—I samples showing an overall independence of AR on M.W. of their samples. On the basis of a fair agreement of micrograph of Figure (5.1 with those of Hashimoto's (1977) and a similar order of rubber block weight fraction it is inferred that a thick domain- boundary interphase exists in this sample. It can be concluded, therefore, that the structure of this particular block copolymer conforms to assumptions in theory of spherically symmetric rubbery domains wiflilow surface coverage uniformly dispersed in a thermoplastic matrix. 6.4 The Modified Weissenberg Rheogoniometer The steady, dynamic, and transient material functions such as shear viscosity, transient, and relaxation stresses and dynamic Viscoelastic functions were measured over a range of shear rates, frequencies, and time with a modified Weissenberg Rheogoniometer, HRG (Model R-16). The modification involved the removal of the axial 80 force servo system and the LVDT transducers and replaced by a dynamic piezoelectric load cell and a charge amplifier. This, along with the utilization of a stiff torsion bar (KT = 5.8492 x 105 dyn cm/.001“ deflection)similar to those employed by Meissner (1972) were made to increase axial and torsional stiffness and thereby diminish unwanted motion in the platen assembly especially during dynamic and Stress growth measurements. Figure 6.3 is a schematic of the internal structure of the NRG. A detailed description and operating procedure will not be given here as they have been reported by various authors and more recently by Cross (1983) on the MRG used in this study. A torque in the torque bar is measured with a linear variable displacement transducer, LVDT. The output voltage is sent through an amplification and low frequency filter units and is recorded on the torsion transducer meter. In event that stress histories are required, the filtered output voltage are recorded with a Honeywell Visicorder that records transient events on photographic paper. An additional clam-shell electric oven was constructed for this equipment to accommodate a Mooney platen of Diameter, D = 10 cm Two types of plate arrangements were utilized in this study are shown in Figure 6.4. 1. The cone—and-plate platen with cone angles of 60 = .552° and 1.982° and D = 7.5 cm and 5 cm respectively. 2. Combined cylindrical and cone and plate platen (Mooney) with 60 - 0.933 Outer Cylinder diameter, 00 = 10.0l cm, 81 I 1 Torque transducer Torque bar _ c Air bearing———T- Constant Temperature____;u’--1 Upper platen oven ‘ I I I I I L - I I Clutch and L—‘--J L---- Lower platen brake Oscillation drive position E 2 Lower platen holder Worm Top bearing I 1 1 3? I s a: Drive box housing I I I I 1 I T\‘\~Linear ball races i1__.Bottom bearing H II 1 I II II Figure 6.3.--weissenberg Rheogoniometer internal (Sangamo Controls Ltd.). 82 L ' J - i I R 1 90 I Li. . ] A Figure 6.4a.--Cone and plate platen. I :0: ml 1 ‘2 —-—a—. O Figure 6.4b.--Combined cylindrical and cone and plate platen (Mooney type). 83 inner cylinder diameter Di = 9.8195 cm. and cylinder height = 2.533cm. Here, the inner cylinder is formed of a conical platen at the bottom and cylindrical side, the diameter of which is accurately machined to allow a radial gap equal to the gap at the edge of the cone and plate of the platen. This ensures a uniform rate of shear throughout the sample. Values for steady and transient shear stresses can be calcu- lated from the torque in this arrangement by noting that for "Couette" cylindrical platens, the tangential shear stresses arising in the gap is given by For coneJand-plate 3.9: xy 2nR S (6.2) where R is the platen radius, h the cylindrical height, and 82 and 9a are the torques developed in "Couette" cylindrical and cone-and—plate platens respectively. Since the shear rate is uniform throughout the gap The total torque = 2(1 + 6h/D) (6.3) The cone-and-plate platens were utilized to collect steady shear and transient shear stress data with a steady shear rate range of 0.005 84 sec'1 to 0.1 sec—1 at 130°C or lower. At 150°C the range improved to v :_0.3 sec-1. Beyond these shear rates ranges shear instabilities were noticed and this will be discussed fully in the experimental section. The Mooney platen was useful in extending shear viscosity data up to v = 3 sec-1. Beyond this an associated error of 9-12% was noted in the viscosity of the calibration fluid (ASTM standard) of n(T = 25°C) = 742.1 poise. This error is attributed to inertial effects and non-uniform shear regime in the gap commonly associated to large size platens performing at large shear rates (Walters, 1970). Due to the limitation of the amount of sample tested with the Mooney platen were limited to the range 0.1 < I < 3 sec-1. The transient and steady first normal stress difference are important material functions normally collected with the NRC. The transient normal stress data of various polymeric melts manifests strong overshoots and sometimes double peaks (Huang, 1976) before attaining a steady state with time. Unfortunately, at the time of this study, the WRG was equipped with a dynamic piezoelectric load cell that registers transient events, but returns to the null state when the steady state is attained. In the light of these no reliable normal force data were collected for the sample. In this work signals from the piezoelectric load cell were displayed on an oscilloscope and utilized in attaining the exact required gap separation distance beween the platens. This was especially useful when using the Mooney platen since it is impossible to see the inner cylinder just touching the outer cup for gap setting purposes. 85 6.5 Sample Loading and Temperature Control In measuring the material functions of SB block copolymer melt, the residence time of the material should be kept very short in order to minimize oxidative reaction in'Uuapolybutadiene phase. 0n the other hand, due to the long relaxation times of polymer melts, rather long waiting periods are required to attain the gap setting and equilibration of the sample to a stress-free initial state. To shorten this period and insure an initial equilibrated uniform distribution of the domains, premolded samples by way of solvent cast films are helpful, with dimensions which fit the cone- and-plate geometry of the test gap. Since the melt temperature is known to strongly affect sample morphology care was taken not to introduce temperature inversions by using the procedure described below. Without setting the gap, the platens are heated to a temperature of 5°C below the desired temperature in about 1% hours. At this point a nitrogen purge of a to 1 lb. pressure is bled into the heated chamber until the desired temperature is attained. It was predetermined that an N2 pressure less than l.5lb. does not affect gap separation nor the torsion readings. After 5 minutescxiattaining desired temperature, the gap between the platens was then set, primarily by the use of the normal force measuring system. Then the oven was opened and the sam- ple is quickly transferred from the evacuated dessicator used in storing'Uwa sample to the plate making sure that no air bubbles were trapped between. The head was then brought down and excess melt cleaned off with a blade and the thermal chamber closed again. 86 The sample was then allowed to heat up to the desired temperature, a procedure that took 45 to 60 minutes. The temperature controller maintained the plate temperature to within i 2°C. 6.6 Rheometric Testing 6.6.1 OscillatorygShear Experiments The measurements of visoelastic properties of poly(styrene-b- butadiene) block copolymer melt were carried out with the cone-and- plate platens at T = 130°C, 150°C, and 175°C. The strain applied to the sample by the oscillation of the bottom plate, causes the oscillation of the top cone. Oscillatory displacements are transformed intoeulelectrical potential by the LVDT. It is then amplified and recorded on the Visicorder. The strain sinusoidal input wave is also recorded on the Visicorder. A phase shift and the amplitude ratio are determined from these two waveforms to obtain the linear Viscoelastic functions as O n'(u>)= —’f=Y—— sin <1 YO So G'(m)=-—5xe cos 0 YO where 0 and 52y‘/io anathe phase shift and the amplitude ratio respectively. 87 An applied strain amplitude range of 0.1 - 0.2 gave no dis- trotions of sinusirdal waveforms in data and this was taken as the linear viscoelastic range. For high sensitivity and small sample size, the cone-and-plate platen of D = 5 cm was mainly used in oscillatory testing. Testing was carried out with the same sample giving from low to high frequency of oscillation. A waiting time of 30-45 minutes between testings was implemented. Oscillatory testing at 124°C with strain amplitude maintained at 0.15 resulted in nonsinusoidal torsion waveforms as shown in Figure 6.5. Such highly non-linear oscillatory behavior have been reported by Ghijsels and Raadsen and a triblock sample and is a peculiar feature with structurizing dispersed systems. 6.6.2 Steady Simple Shear Experiments Low drifts were noted in the torsion head transducer meter range of 0.25 x 10.3 in and 1 x 10‘3 in the gap. Therefore, a shift torsion bar (KT = 5.8492 x 105 dynes cm/.001 in.) is utilized as it gives the highest sensitivity at the transducer range setting of 2.5 x 10'3in. Such choice was made to restrict the movement of the torsion head to a minimum aiding transient measurement with the chosen platen diameter and the anticipated value of the steady viscosity of the sample. A steady shear rate range of 0.005 to 3 sec-1 was attainable with the instrument using both the cone and plate and the Mooney 1 platens. Data were obtained in the range 0.005 to 0.l sec- with a waiting time between measurements of 30 minutes and next with one hour. 88 .mH.o co mnzpraEm :wmcpm pm mmmcpm camsm xcopm—Fwomo co Ecocw>m3 PmcwomzcwmucoZ11.m.o «camel NHH.o mH.o wovmfi u u u l— o ’3 >. page? cowpm___omo “amuse cowmcoe 89 No appreciable difference in data was noted and thus the former waiting period was implemented. An associated error of 7-10% in the material functions occurred in this range. Beyond this range and at a tempera- ture of 130°C or less a variation of 12-20% was noted in the stress readings for different runs under the same conditions. Upon closer studies it was observed that shear instabilities, e.g., stress frac- ture developed in the material as can be determined in Figure 6.6a and 6.6b. Figure 6.6a illustrates the situation where the material is extruded out of the gap after a shearing time of 8 minutes. In 1 and Figure 6.6b the appearance of the material at y = 0.096 sec- 0.43 sec'1 are compared for quenched samples which experienced similar shearing times. Non-uniform shear profile is likely to devel0p in the sample at i = 0.43 sec"1 resulting in faulty stress readings. The Mooney platens have a potential range of 0.1 < I < 10 sec'1 as seen in Holden's data. The major advantage of this platen is that the sample is prevented from leaving the shearing gap by the guard ring. Also very little area of the material is exposed to the air minimizing errors due to oxidative degradation. However, the bulk (D = 10 cm) of this platen tends to increase the inertia head leading to inacuracies mainly in transient and oscillatory measurements. As can be seen in the viscosity flow curves (presented in Chapter 7) no appreciable error is incurred using this platen at the shear rates prescribed as data extends smoothly iflxmi low to moderate shear region, i.e., (0.1 < v < 0.4). The effect of inertia is, however, seen in transient measurements as will be shown shortly. For steady shear 90 Figure 6.6a.--Picture showing test material extruding from gap after a shearing for 8 mins; 7 = 0. 3 sec'l. Shear insta- bility is due to stress fracture. T = 130°C. \ ~ “'31 ( 1 Figure 6.6b.--Quenched sheared materials after a shearing time of Left hand specimen sheared at Y = 0.096 8 min. Right hand specimen sheared at T = 0.43 sec-1. sec-1. T = 130°C. 91 viscosity results an associated error of 5-7% was noted using the Mooney platen. Using fresh samples, stress growth experiments were con- ducted with the two platens. After the temperature of the material has stabilized in the gap, the clutch system was quickly engaged after the motor has been running for at least 5 minutes. Stress transients were recorded on the Visicorder that was calibrated with the steady state stress value obtained from the torsion head transducer meter. The time dependent stress is normalized with the steady state value. Using the Mooney platen the effect of inertia on transient measurements can be seen in Figure 6.7. An overshoot in the stress build up does not occur until at v = 0.914 sec'l. This is in sharp contrast with results using the cone-and-plate platen at the same temperature, which shows an overshoot at shear rates as low as 0.027 sec-1. It is generally observed that overshoot occurs in stress-growth at high shear rates. It denotes the point at which the material experiences a maximum strain. Transient measurements are also affected by the cone angle of the cone-and-plate arrangement. Theoretically, the assumption the cone angle, 60 is to be chosen such that the assumption tan 00 ~ 00 is valid. This insures the existence of a constant shear rate throughout the melt. Meissner and Huang noted systematic differences in transient shear stress and normal stress measurements as a func— tion of cone angle used. However, Graessley et al. (1977) found no difference in stress growth measurements for 1°, 2°, and 4° cone. 92 .cmpwra chooz one mcwm: goomfl pm m1m mo :owpuczw gpzocm mmmcpm cmmcm11.s.o mesmwe 3 NH 3 w u m e N o omud ezwo.1+ N6 m.o 04 N." 93 In this study on comparing measurements as a function of the cone angle of 0.552° and 1.982° gave a variation of data of 1.8% which is well within the experimental error. It is thus presumed that the choice of 00 : 2° introduces no significant error in the transient measurements. The stress relaxation after cessation of shear was also collected using the Visicorder on samples used in stress growth tests. The results of these experiments will be presented and analyzed in the following chapter. CHAPTER VII RESULTS AND DISCUSSION 7.1 .Introduction The material functions, dynamic viscosity, storage modulus, steady shear viscosity, shear stress growth, and relaxation stress after cessation of shear of a poly(styrene-b-butadiene) with 94.1 wt. % S have been collected as functions of the deformation rate and temperature. These results suggest that there exists a melt transi- tion temperature demarcating the prevalence of two types of block copolymer microstructure. The occurrence of such transition tem- perature or region will be discussed in Section 7.2, using evidence in the experimental results. We will not use the time-temperature superposition principle in reducing data since such two-phase struc- ture in block copolymer melt have been established (Chung and Gale, 1976; Gounlock and Porter, 1977). In Section 7.3 the rheological results showing the effect of deformation on the microstructure above transition temperature shall be presented and discussed with the view to understanding the underlying microstructure. Next, the rheological results below transition temperature is presented and discussed again with a view to verifying the block copolymer microstructure. In order to test the transient network model developed in Chapter III, two material constants are estimated using 94 95 the linear viscoelastic data. Other model parameters shall be com- puted by fitting model predictions with these functions. The model then will be used to predict the steady and stress growth flow behavior of the block copolymer at T below the transition. It is necessary to restate here that our major focus is on the rheologi- cal behavior of the block c0polymer at low deformations. This region yields the most differentiating features of block copolymers with respect to their homopolymer and random block copolymer counterparts; it also plays a crucial role in evaluating a network model based on a more realistic chain statistic. 7.2 Phase Transition Temperature The dynamic viscoelastic properties of the block copolymer sample are shown in Figure 7.1 at T of 130°C to 175°C. The repro— ducibility of these results is good, 5.2% at w < 0.6 sec'1 and fair, 1 with 2° cone angle and the stiff torsion 7-9% over 0.8 < w < 3 sec- bar. As usual, increase in temperature tends to decrease the moduli. At 150°C the dynamic viscosity levels off at about 0.1 see”1 but stor- age modulus as a function of the frequency shows a slope of 1.3 on the log-log scale. On the whole, such behavior is similar to those exhibited by homopolymers where a single phase microstructure is known to exist. At 130°C or lower, the dynamic viscosity does not level off at the lowest frequency tested and a larger deviation of the slope of the dynamic storage modulus vs. frequency from 2 is noted. Next, we evaluate the two temperature regimes for homopolymeric character by applying the Cox-Merz rule defined earlier on the viscoelastic 96 .mmczumcoaeop msowca> pa m1m we zocmzcwcm .m> m:_:coe mmmcoum wen zpwmoomw> quach11.H.m wcsmwe S 2 TS A o _ _ NS 0 S a 41 1 ) «2 2m 6 / .U S O I e 0 ) my 10 d w. - .‘ a G \\ ( Q . S1 .. . m 4m”: 0 O; O ”is o . UoOmH I O 8.02 o . S e 02 97 properties. These results are shown in Figures 7.2, 7.3, and 7.3a. At 150°C as given in Figure 7.2, the complex viscosity is found to be greater than the steady shear viscosity especially at v > 0.1 sec-1. 1 However, at v < .1 sec- both functions not only level off, but appear to be approaching each other. The steady shear results show a limit- 5P. This compares with a homo- ing zero shear viscosity of 1.4 x 10 polymeric PS having MW = 259,000 Mw/MN = 2.35 at T = 200°C with no = 4.25 x 105 P (Mendelson, 1980). On the whole the deviation from Cox-Merz rule follow a similar trend often shown by homopolymers and random copolymers. A dissimilar deviation from the Cox-Merz rule is found when the complex viscosity, the dynamic viscosity and steady shear viscosity results at 130°C are compared as reported in Figure 7.4 and 7.3a. In Figure 7.3 the largest deviation of the two functions appear at I < 0.1 sec-1. Both functions are sensitive to the deforma- tion rate at the low deformation rate region suggesting a more complex microstructure controlling the viscoelastic response. At w ~ 0.01 sec—1, the complex viscosity appears to be levelling off even though more data (at w ~ 10'3 sec-1) are needed to confirm this assertion. If this is the case, these results suggest the occurrence of a network structure sensitive to the imposed strain history of the material. Figure 7.3a shows the dynamic viscosity, n' to be significantly sen- sitive to the frequency equivalent to the shear rate range 0.05 < Y < 0.3 sec'1 where a so-called "equilibrium" shear viscosity is attained. At w < 1.5 sec-1, n' values are higher than those of n by 18% or less. The strong dependence of n' on w seems to lessen at 98 .>_m>wpowammg .mp8; camsm vcm zucmscmcw +0 cowpuczc a mo ooomfi pa mim co zpwmoomv> hummpm new xmraeouii.m.m mesmwe Afiiommv >.3 {OH OCH HIGH NIOH d . . voH (asiod) |¥u| oofi .x_m>wuomamwc .mpwc cmmzm new socmzcmcc co cowpocze a ma moomfi pm m1m co prmoomw> summpm new meQEOUii.muN orgasm AH1ommv > .3 OOH H1oH NioH 0H 99 _ _ _ OH '1 OH 100 floH .>~m>wpowamms .mpmc cmmgm ucm xocmscmcc co coepocsw m we ooomfi pm mim co haemoumw> anmmpm new ovens» 11.8m.~ oczmwe c» .3 OOH s-oH N-oH . . 4 OH UoOMH H H OH 101 '1. Further data at w < 10'2 sec'1 will be very helpful w - 10'2 sec in establishing whether n' levels off and how this frequency at which this occurs compares with that suggested for |n*|. Since the zero shear limit for 150°C is seen at shear rates comparable to that at which homopolymer n levels off, a single phase microstructure is suggested. For two phase structure, such a limit may be observed only at deformation rates that are order(s) of magni- tude lower. Attainment of Newtonian viscosity at such low shear rates implies the prevalence of a network structure sensitive to the applied strain history, that have been attributed to diblock copolymers (Krauss et al., 1971). The transition temperature region for this diblock sample occurs at 130°C < T < 150°C. This is attributed to a weakening and/or loss of the two-phase structure due to sharp increase in phase miscibility and/or the attainment at or above the transition temperature of an easily disruptible dispersed phase not controlling viscoelastic response and, therefore, leading to NeWtonian behavior at low deformation rates. The narrowness of the transition suggests that chain miscibility is at least the major factor since, in the absence of suchaiphase change, the property changes would be expected to be more gradual. 7.3 Viscoelastic Behavior Above the Transition Temperature Figure 7.4 shows the effect of shear rate on the steady shear viscosity at 150°C. We note here the quick decay of viscosity from Newtonian behavior at higher shear rate. Such behavior is often 102 .UOOmH pm m1m co cowpoczo zpwmoumm> camcm znmmpm11.¢.m wczmwe -smwe s A as Os H o H-0H - 4 OH (asiod) 0 m3 OH 103 associated with high M.W. polydisperse homopolymeric melts. The polydispersity of the continuous PS chains in the sample under study is 1.7. In Figure 7.5 the shear stress growth results are portrayed as normalized values using the constant stress value S as the 1 xy/ss shows points normalization constant. The curve of v = 0.0108 sec- of inflection at 8 mins. and 16 mins. that are not found in the other curves. The associated error observed for this shear rate between forward and backward rotation was 3-5% at t < 2 min, 12-18% at 2 < t < 12 min. and about 5% at larger times. It is judged that this error may be caused by incomplete relaxation of the test sample in the gap. The other results reported in Figure 7.5, as well as the shear stress growth curves at 130°C had associated errors at 5-8%. 0n the whole, the trends in result resemble those of homopolymers. We note, however, the occurrance of significant overshoots at much smaller shear rates in contrast to homopolymeric melts (Graessley et al., 1977). Furthermore, these curves depart from linear visco- elastic behavior even at small times. The extent of this departure may be determined by evaluating a relaxation modulus G0 at different strain rates from the slopes of the normalized growth curves at small times. +/S S G0 = lim xy xy,ss t->0 t The values computed for GO ranged from 0.28 to 0.52 over the range of shear rates studied. Figure 7.6 shows normalized stress relaxation functions at 150°C. 104 ooomfi pm m1m co cowpoczc zygocm mmmcpm camsm umNPPmE20211.m.N mczmwe Aommv H mm mm «N om 0H NH w a c _ _ a q H _ J A I 0 rs 0 iv L m0 9 O 43 moflo. - L..1 0 owo.o memo.o 1 we Hmfio.o n s x xs ss/KXS/ 105 .. A83 5 .0003 a.“ i we cowpmxmpmc mmmcpm cmmsmii .0.“ mczmwe o.¢~ o.oN o.oH o.m~ o.w o.¢ o a _ d ammo. memo. ono.o 1.N.o 1: s X nA. / S VA .1 as m + m.o 106 These results showed a decay of stress relaxation to zero similar to those of homopolymers. It is concluded that an entanglement microstructure of the continuous PS chains appears to influence the block copolymer melt at 150°C or higher; however, these results do not exclude the existence of domains above this transition temperature since easily disruptable domains not controlling the viscous response might yield similar results. The range of temperatures where a two-phase structure mani- fests in the material has been established at T < 150°C. In keeping with the objective of this study, viscoelastic results of 150°C will not be compared with the transient network model. In the next sec- tion, the material functions of 130°C shall be presented and compared with the model having a nonzero "a." In this analysis the tangential shear stress shall be normalized by the constant G0 = nokT, the modu- lus of rubber elasticity, while the shear rate, v and present time, t are normalized by a single relaxation time 10 ( = l/BO). It is worth emphasizing here that our interest lies in the low deformation rate region and we seek to predict the material viscoelastic behavior at such a range. Since linear dynamic functions were not obtained at 124°C, it is not possible to predict other material functions at this temperature using our procedure. In this model evaluation we will deal mostly with normalized quantities, i.e., the normalized viscosity, normalized shear rate, in line with definitions given in Chapter V. 107 7.4 Viscoelastic Behavior Below the Transition Temperature The complex viscosity as a function of frequency, shown in Figure 7.3 has a slope, d log |n*|/dlogw(at w < 0.1 sech) equal to -0.26 as compared to a value of -0.5 obtained by Ghijsels and Raadson for SBS triblocks. Table 7.1 further illustrates the results of the slopes of n and n' vs. v and m respectively, (see Figure 7.3a). Table 7.1.--Phase separated block copolymer melt properties Sample MN(x10'3) TOC dlogn'ldlogwl)<1 dlogn/dlogi/Y<1 535a 11-56-11 150 -0.61 -0.68 535b 22-50-22 170 -0.66 -0.66 585b ' 14-70-14 170 -0.35 -- SBC 232-10 130 -0.43 -0.38 aData of Ghijsels and Raadsen (1980). bData of Arnold and Meier (1970). cThis work. These slopes indicate that the triblocks have more strength than the diblocks (even at higher melt temperatures). Even though the SB diblock has an MIT an order of magnitude higher than the SBS triblock, from these slopes the network structure of the triblocks are stronger. The results of steady shear viscosity of S-B melt sample as a function of shear rate at 130°C and 124°C are reported in Figures 7.7 and 7.8, respectively. The upturn in viscosity occurs at about 108 HoH Afi1ommv > OOH H1OH NioH .ooomH pm .m1m co mam; emmcm co cowpoczw a we mnemoomw> Lamsm Avmmpm mghii.N.m mesmwe 1oH d _ _ ommm.H u o memow_a mum-Q new mcou o O ommm.o u o mcmmeQ Amcooz mcvm: . «OH (aslod) 0 0H 109 HoH .UO¢NH pm mim co mow; cmmgm mo cowpoczm a ma zpmeUmw> camsm aummpm osp11.m.m mczmwe -ooH Afliommv > T2 FE TS. d OH OH (astod) u coH 110 1 at both temperatures similar to an SBS melt at 150°C 0.1 sec- (Ghijsels and Raadsen). About shear rates of 0(1) the viscosity is no longer strongly dependent on the shear rate, but thereafter the material seems to approach the power law region. Upon comparing these curves with the high temperature curves (T > 150°C) we see that the low shear rate response is that of a weak three dimensional micel- lar network in which the polybutadiene domains acting as junction sites solely influences the viscoelastic response. Further evidence of the effect of two-phase microstructure can be seen in the shear stress growth curve5(fi°Figures 7.9 and 7.10 collected at 130°C. At small times, these curves exhibit higher transient shear stress with lower shear rates, than the corresponding curves at 150°C. It is further observed that the magnitude of the overshoot from the steady state level is higher (0.18) at v = 0.0272 sec"1 than at i = 0.043 sec'1 (0.12)--a feature also present in the SBS data. On comparing these curves with the high temperature counter- part (Figure 7.5), it is clearly evident that a more detailed micro- structure behavior is found in such transient flows and more effort should be applied in this area for a better understanding of the microstructure mobility than at steady state conditions. No stress growth responses were obtained at 124°C due to the limitation of_the amount of sample. The next curves (Figures 7.11 to 7.13) shows the stress relaxation functions at 130°C and 124°C. Here in contrast to the findings on SBS data which manifests residual stresses, these 111 mm mm uoomH pm mim mo cowpocze cwzoem mmmgpm cwwcm nmNPFQEL0211.m.N mezmwe vm om 0H NH mmooo.o ano.o mmmo.o n + m.o v.0 m.o m.o o.H SS/KxS/KXS ‘1“ 112 (\I (‘0 mm 00 OMH Hm mum Lb Cowpocsuv LHBOLG mmmme mecm U¢NZ.©ELOZII.O~.N stwwn— 0N 0m 3 6mm 0 NH 0 _} 25.0 n s m0 0.0 0.0 0.0 04 NA 113 .ooomH pm 01m 00 we?“ co cowpoczw a ma cowwmxmpmc mmmcpm cwmnm nmeFmELozii.HH.N 0030?; 140' Aommvp 0H NH 0 a 0 0 58.0 a 380.0 1 s 5 .0 m .V / S VA tn / w 0 m. 8'0 114 0H .ooOmH Pm mum Lb 9:5. *0 :0.5.U::.+ 0 mm Cowymxwrw.» mmmLHm wazm UmermEe—OZII.NH.N 9:09.... Aommvp 0H 0H NH 0H 0 0 0 N — a . . _ v _ . . 115 .00¢NH “e 01m ee mafia we eewpecee e we :ewpexepec mmeepm Leesm eeNWFeELe211.mH.N mesa-r Aeemvp em era 0.8 0.3 0.2 08 or. e .. .- , q _ _ e e 2e memes 1: 58.0 n e 1: 5 X. ”A / S X »A / 0 es 1e.e es 116 functions at all shear rates decay to zero, but at a much slower rate than those of 150°C. At higher shear rates (i > 1 sec-1), polydispersity of the continuous phase in our sample makes it difficult to determine.whether domain flow or entanglement disruption in the continuous phase con- trol the viscoelastic response. Whether domains are completely dis- rupted by shear deformation and the point to which this occurs may be difficult to establish with rheometry alone. This may be made possible by utilizing electron microscopy with deformed samples as was performed in solid elasticity (Aggarwal et al., 1969). This is outside the scope of this study. 7.4.1 Estimation of Model Parameters The non-Gaussian transient network model presented in Chapter III assumes that the continuous soft phase of the block copolymer is composed of the "most probable" network segment with N sub-units. This demands the knowledge of a single relaxation time 10, that is associated with the rate coefficient, 80, (10 = 1/80) of the destruction rate process and the modulus of rubber elasticity, 60’ (G = nokT). In reality, in any polymer matrix, there is a distribution of N and thus multiple relaxation times obtainable from the fluid relaxation spectrum which is often constructed from functions of linear viscoelasticity, G'(w), 6"(w), and G°(t). It is worthwhile to emphasize that G0 and A0 are not to be considered as adjustable parameters in the model. 117 Even in homopolymer rheology the use of a single relaxation time in viscoelastic models can only predict data in a restricted range. Generally, a large relaxation time characterizes long time behavior and is applicable with low deformation rates predictions while a small relaxation time predicts higher order deformation rate range. Typical dynamic shear moduli of narrow M.W. distribution samples display two sets of relaxation times corresponding to two relaxation mechanisms separated in the time scale. One set of relaxa- tion times associated with the transition in the high frequency region; another set associated with the entangelment slippage in the low frequency region which appear as a peak of G"(w). A character- istic relaxation time associated with long-range motions of homopoly- mers is estimated by the inverse of the frequency at which the peak of the loss modulus, G"(w) occurs (Onogi et al., 1970). How- ever, in polydisperse samples there is often an overlap between these sets of relaxations so that the peak in G" appears as a plateau. Further, the slope of 0" vs. w is close to unity on a logarithmic scale for homopolymers. Gouinlock and Porter have identified that the departure from 1 of this slope in block copolymer melts is due to the domain morphology. Ghijsels and Raadsen also found the presence of maximum in the loss factor tan 6 (= G"/G') and related this with domain activity. These points were considered as one of the criteria in determining GO and A0. The other criterion is based on In 118 the point where the upturn of viscosity occurs in the experimental steady shear viscosity. Such an upturn also occurs in the predicted curves based on the former criteria, but they were plotted as a function of 10?. By comparing these points 10 can be evaluated. Figure 7.14 shows the results of G"(w) and loss factor as a function of frequency. The deviation of the slope of G" vs w from 1 is not very discernible but the loss factor shows a pronounced transition at 0.25 sec-1. From this we obtained the material con- stants shown in Table 7.2. Also using the refining criteria a second set of relaxation times are evaluated and listed in Table 7.2. Table 7.2.--Material constants from experimental data Method 1 Method 2 10(sec) 4 1.25 d nes 4b 5 G (—L—) 6.8 x 10 2.16 x 10 o 2 cm a _ A0 — 1/00t b : Go - |G*(wt)| Values of the segment repulsion range parameter "a," the destruction rate coefficient "a" and the slip factor a for Method 1 are determined by fitting the data of complex viscosity with Equa- tions (5.26 and 5.27). The result of the best fit with data is given in Figure 7.15. In Method 2 it was necessary to refit the data with 119 tan 6 .ooomH we xeceeeec$ 0e mcewuecee me segue» mme_ use m:_:eee mme411.¢fl.w eczmwe A OH eeH H es 0H 0.0r 0.H1 ¢.~i ZwD/UKP (m)119 0.H T m.H 0H 1‘1, 120 .'.asa. n. ‘ . - 5...... 10 ._ I e = 0.005 ' a = 0.45 I g = 0.03 A = 4 sec. ' o 1.0.. )- T I LC; 0.1 1 J I I 141 l l l l J 4411 j 1 J 0.1 01A 1.0 0 Figure 7.15.--Evaluation of model parameters using linear visco- elastic functions (dimensionless). o = data, ——-—- = model fit Equations (5.26) and (5.27). T - 130°C. 121 the new set of constants and the results are shown in Figure 7.16. These results are least sensitive tO'Hueparameter "e", the range 0.001 < e < 0.007 gave practically the same results. This parameter is best ascertained with strong flows, e.g., in uniaxial transient extensional flows. 7.4.2 Experimental Evaluation of the Transient Network Model Without any further adjustments in the parameters, steady and transient shear results are predicted by using Equation (5.17) and portrayed on the accompanying plots as a normalized viscosity (fi (n/Golo)) as a function of normalized shear rate, (A0?) and normalized transient shear stress (5:y/5 ) as a function of xy/ss normalized time (t/AO), respectively. 7.4.3 Steady State Predictions The model predicts correctly the overall trends of the steady shear data as shown in Figure 7.17 and 7.18. In Figure 7.17 the quantitative agreement between experimental results and theory is poor to fair in the range 0.02 < i < 0.18 where a 40 - 0% deviation is noted. The theoretical prediction of the range 0.18 < § < 12 is satisfactory with about 5% derivation. 0n the other hand, using Method 2 having the same order of magnitude of the relaxation time as in 1 improved results signifi- cantly at the range of interest (see Figure 7.18). In the range 0.006 < i < 0.1, the theoretical prediction of results is excellent having under 3% deviation. At moderate dimensionless shear rates of 122 .uoomH u H .AoN.0V use AON.00 mcewpeeem .pww Feces u .epme u 0 .Ammepcewmcesvev mcewpecey UTHmePeeemw> Leecwp meme: meeHeEecee Feces we seepe:_e>011.0fi.s 8230?; II r< omm 0N.H 00.0 000. 00.0 n n u 0» OJ re L 0H erI 123 .uoomH u H .868 e 1 OK .me.e 1 m .me.e u w .mee.e 1 6 Fence 026256: pcewmcecu ;a_3 xnwmeemw> geese seeeum 0e cemwceeee011.NH.N ecemwe 0H 0.H 004 H.o 1 A _ _ 4 _ H.0 0.H 0H 124 0a .emm 0N.H u e& .00.0 n m .00.0 n 0 .000.0 u 6 Fence saw: xuwmeemw> geese meeepm we cemwceeee011.0fi.u ecemwe 0.H . H.0 H0.0 o 6 logo. 0H 125 0.1 to 1, the prediction is fair to poor with 3-36% deviation and unsatisfactory at I > 1 with 40% deviation. The slope of viscosity as a function of shear rate at v < 0.1 sec"1 is predicted very accurately. It is concluded that the loss factor cannot serve as a guide in obtaining a characteristic relaxation time. At the large shear rate range, it is unreasonable to expect a good fit in the light of the polydispersity of the PS phase (MW/Mn = 1.7). 7.4.4 Transient Predictions Comparison of the model predictions with the data for stress growth are given in Figures 7.19 to 7.24 at low shear rates using the two procedures. Here the agreement between data and theory is rather fair, especially if we remember that all the parameters were determined from data of small amplitude oscillatory shear flow only. On the whole, the model prediction with 10 = 125 sec is good at the lowest shear rates (0 - 15% deviation) and excellent at strains less than 9.001 (under 3% deviation). On the other hand, the high relaxation time model appears superior at higher shear rates for all models significant deviations occur at intermediate times. A weakness in the model is its failure to show an overshoot at low shear rates. Such overshoots are shown at higher shear rates, as illustrated in Figure 7.25 on page 132. The positions of the overshoot, tC can be correlated most directly with the total strain as many workers have noted previously. Figure 7.25 shows the strain at stress peak as a function of shear rate for both data and model predictions. Overshoots are predicted by the model only eFmeeE II 233 Iol .80 u w .000 .1. e .03 v n K .302: spy; 06000 we mppemmc guzogm mmmgpm we :emwgeeee011.0fi.m eczmwe 126 NNN0.0 u 0 A O D n O O \s 0.0 0.0 0.H 127 1.0 _ 10? = 0.00856 0.4 one .5 N ._ A 1.— t/AO Figure 7.20.--Comparison of stress growth at 130°C with model A = 1.25; a = 0.55, E = 0.05 0 = dat8;____ model. 128 .eem e u e« we Feces Law; uoomfl 0e cpzecm memepm we cemweeeee011.HN.m ecemwe N.0 0.0 0.0 0.0 0.H ox\e 01 0 0 0 v m N H o _ q a _ _ 4 ~ 7 1 4 . J . 1 s 4 a i o O J 000H.o n 0 K e o D D c 1. HI I} 0 O a O a O 1 - o L .N.H ss/KxS/A‘xS 129 O .0N.H n K %e Feces new; goomfi we £02000 mmeepm we cemweee5001i.NN.m ecemwe 0 mm em a em 8. Se 2 e e N u H d 1 . 4 q d - O emee 1 + A 0 O 0 o 0 o 0 0 0 O 0 130 1.2 F Figure 7.23.--Comparison of stress growth at 130°C with model of 10 = 4 sec. 131 0H 0 .emm 0N.H u < we Feces new: ooomH 0e gpzeem emerge we cemwceeee01-.¢N.n mcemwe 0H NH 00 0 0 w CI emme.e 1 +60 6 0 b O fix *8 ss/KXS/ 132 OH .111 .000.0 n o .00.0 n 0 .00.0 1 m .weuee .eueu n 0 .0oomH we mm>L=e geese pcewmcegu cw peegmge>e mmecpw ecu we :wecpmii.0m.n ecemwe eem 0 e; r- 0 no _ _ . _ . 0.0 0H 133 at v > 0.25 sec-1. The predicted peaks occur at strains insensitive to the shear rate and are determined by the slip factor a. A constant value for this strain of about three has been reported experimentally for a homopolymeric melt (Osaki et al., 1976). Graessley et al. (1977) have studied this quantity at low shear rates with homopolymer samples and indicated that insensitivity of strain at stress peak, to shear rate is associated with materials that possess a broad relaxation spectrum. It is noted that the foregoing feature and the fact that the magnitudes of the overshoot for transient stresses at smaller shear rates are larger than those at large shear rates (which is not pre- dicted by this model) presents a severe test for viscoelastic models. This will have to be addressed with only £33; relaxation time if the exact physics of two—phase microstructure mobility is to be compre- hended. CHAPTER VIII CONCLUSION AND RECOMMENDATION 8.1 Conclusion A new kinetic network model has been developed and evaluated for the rheology of block copolymer melts and polymer composites with spherical microdomains. This model involves in addition to the readily determined relaxation time A and modulus GO, three parameters: ”a" describes the range of repuslion between segments of matrix attached to spherical domains, "e" describes the dependence of junc- tion destruction rate on the conformation of the continuous random phase and (g) accounts for a slip between the fluid and the network junctions. The m0del hsused to compute the material functions in uniaxial extension, simple shear and small-amplitude oscillatory shear flows. Experimental data on elongation are obtained from the literature while datacrishear flows are obtained in this work. In uniaxial extension, the model predicts the Trouton viscos- ity at normalized strain rates, I of 0(1) if spherical domains are absent (a = 0). This is in good accord with data of Mundstedt and Laun (1978). If spherical domains are present (i.e., a f o), the model predicts a non-constant elongational viscosity at the low strain rates, but a smaller maximum viscosity at higher strain rates. Com- parison of these calculations with data of ABS melt (Mundstedt) 134 135 reveals that the repulsion measure ”a” determines the apparent yield stress observed at low elongation rates. The destruction rate para- meter "8" determines the level of the maximum elongational viscosity at steady state as well as the stress overshoot observed at higher rates in stress growth experiments. However, no data for elongational flows at large strain rates are available to evaluate the model suit- ability in this region. The viscoelastic properties of a diblock copolymer, poly- (styrene-b-butadiene) of high thermoplastic content have been studied experimentally in this work. Theinaterial is composed of uniform spherical domains of polybutadiene randomly dispersed in a poly- styrene matrix as confirmed by electron microscopy on solvent cast samples. The melt for rheological study was obtained from carefully annealed solvent cast samples (toluene as solvent) leading to an associated error of 7 to 10% in material functions at low shear rates. An associated error of 12—20% have been reported by Ghijsels and Raadsen in melts starting from crumbs in this region. Microphase separations appear to start as the temperature is lowered from 150°C. At 150°C or above the material exhibits Newtonian behavior in the steady shear viscosity andtimrcomplex viscosity at low deformation rates and appears to obey the Cox-Merz rule. At 130°C or below the complex viscosity is higher than the corresponding steady viscosity, except at very low strain rates (i < 0.05). At 130°C and 124°C a significant upturn of steady viscosity occurs for shear rates lower than 0.1 sec-1, similar to the SBS melt; 136 however, the slope d log nld log 1 1+0 is much less for the present SB melt. In contrast with homopolymers and random copolymers, a significant transient stress overshoot is observed in the shear growth experiments at shear rates as low as 0.02 sec-1. It is further noted that the height of this overshoot diminishes with increasing shear rates. Contrary to the SBS data, no residual shear stresses are observed in the SB data in shear stress relaxation experiments confirming the assertion that only apparent yield stresses are exhibited by the SB melt. All model parameters have been found by fitting data of oscillatory shear experiments, using two procedures to obtain the characteristic relaxation time. The overall trends in the data have been predicted very well, in the range of interest. Quantitatively, the predicted shear viscosity is very sensitive to the choice of the single relaxation time at the low shear rate range. The model also fails to show an overshoot in stress growth at shear rates less than 0.1 sec-1. These deficiencies are largely due to assuming that only a single relaxation time controls the entire material viscoelastic behavior. 8.2 Recommendations forFUrther Study The stress constitutive equation presented in equation (3.12) is written for a single relaxation time. To take into account the distribution in N, especially in polydisperse samples, we must allow for multiple relaxation times. Since in the network theory no inter- chain correlation is taken into account, each active network segment 137 therefore contributes to the stress additively. The overall stress then becomes: III/3 (9.1) where gireplaces S, N becomes Ni’ and f becomes f1 in equation 3.12. The consequence of this in the specific stress relations is that Go and 10(- l/BO) are replaced by Gi and 4i respectively obtainable using the material relaxation spectrum H(Ai) through these relations (Phan Thien and Tanner, 1978). I H(i)xdx A. = (9.2) 1 Hum H(A)dx G1 = A (9.3) Here the relaxation spectrum is subdivided into intervals, such that each interval is a wedge spectrum to facilitate the numerical proce- dure. The relaxation spectrum can be computed from the linear vis— coelastic data G'(w), G"(w) and G(t) by the standard method (Ferry, 1961). The long-time behavior of block copolymers is of utmost significance in gaining the Optimum relaxation spectrum, thus it is necessary to collect G'(w) and G"(w) data at frequencies as low as 10'4 5-1. Such ranges are achievable by using the cone-and-plate platen with several cone angles, which were unavailable at the time of experimentation. 138 For a complete knowledge of melt rheological behavior of block c0polymers, steady and transient normal stress data is highly needed. Chung and Gale and Kraus and coworkers (1971) have associated the material exuding from the shearing gap at low deforma- tion rates with high elasticity developed in the material, but did not report any normal stress data. In the WRG the shearing gap is significantly influenced by the lack of vertical stiffness of the apparatus. This lack of stiffness affects both steady and transient normal stress response measurements of molten polymers (Huang, 1976). Modifications to correct for this problem were given by Hansen (1974) and is recommended for this equipment. The use of Mooney platens of D<:5 cm along with a steady piezoelectric load cell is further suggested. Curtis and Bird (1981) have presented a reptation theory for melts starting from the general phase space formalism (Bird et al., 1977). They modeled the macromolecules as Kramers freely chain (with N beads and N-l rods of length a) used a nonisotropic version of Stokes law to describe the drag force on a bead as it moves through the melt. The model contains four parameters, the number of beads, N, a drag coefficient r, a link tension coefficient EB and a chain constraint exponent Bc' They report that the model yields no ~ M3'13c. 6+2BC and 1p1,0“M which cmpares well with homopolymeric data if BC~ 0.3 - 0.5. Modelling of polymer molecules as beads joined by elastic or rigid, connectors is attractive for block copolymeric systems 139 with spherical domains. However, the Curtiss-Bird theory does not allow us to compute the relevant chain segment distribution function. A recommended route would be the concept of configuration-dependent molecular mobility tailored by Giesekus (1982). He associated a tensorial drag coefficient 21 with the force, fi’ experienced by an ith bead. This drag tensor-does not depend on the actual configura- tion of the molecule, but only on the average configuration of all the molecules. After some manipulations with the excess stress relation, a configuration tensor 9i can be defined which maps the actual molecular configuration from the equilibrium configuration as o o 1 0 0 .r. > = l 3 < r.- r.> b. —J This tensor may be understood to be a measure of deformation of an elastic continuum, note in a strict sense of a material con- tinuum, but in a statistical sense represents only the configurational states of a polymer chain. With this assumption one may no longer assign individual bi and g to every position vector r1. Instead the whole set of beads (i a 1 . . .; N) can be classified into classes (K = 1, . . . K) with "K beads per unit volume with a common configuration tensor 2K and a drag tensor 5k for each class. The class K = 1 leads to various Lodgean type models with appropriate assumptions on b, but classifi- cation of the total number of structure elements into K classes may 140 encompass systems such as block or graft copolymers. Here only detailed modelling of 2K is required to generate the constitutive equation. In this study the major focus was on spherical domain block copolymer systems, but as shown in Table 2.1 cylindrical and lamellar type systems possess superior rheological properties. 0dani et al. (1977) have studied diffusion, solution, and permeation behavior for a series of inert gases in block copolymer films having these morpho- logies Ihinted that they were excellent models for understanding the i 7 relationship between the morphology and transport properties hetero- geneous polymeric media. The preparative methods of these block copolymers have been much refined by the Dow Chemical Company, Mid- land). It is recommended that rheological and transport studies of block copolymers of higher block composition be undertaken. APPENDICES 141 APPENDIX A UNIAXIAL EXTENSIONAL FLOW--TRANSFORMATIONS AND CALCULATIONS 142 APPENDIX A UNIAXIAL EXTENSIONAL FLOw--TRANSFORMATIONS AND CALCULATIONS The solution of (4-3) using the method of characteristics is given by: t a A 4 f(x ,y , 2,t) = f(xo,yo,zo)expf-]B (xoexp(1‘t’),yoexp(-I‘t'/2), zoexp(-I‘ t'/2))dt'J O t I o I A + .LG(XOe(part),yoexp(‘TtI/2),Zoexpc-IQLI/Z).N)€Xp[- at” ti X B (Xoexpfi‘ tfl)IYoeXp (-’I?I:’//2),Zoexp(-’I§t”/Z))Jdt' (l-A) A two-step change of variables similar to that of Fuller and Leal (1981), but for uniaxial extension is performed on (1-A) to intro- duce definite limits on the integrals. First t' and t" are changed to x' = xoexp(§t') and x” = xoexp(§t") respectively; then x' to 2 using x' = Tx + xexp(-Pt) where T = 1-exp(-§t). Further, y = y(x/x')%, 2 = z(x/x')%, 0' = x"/x to obtain a final expression for f(x,y,z,t) as _ A A f(x’y’ z’t) _ f0 (xexp(-I‘t), yexp(I‘t/2), zexp(’Ft/2)) exp [’If8( [CH 99"- P (’Ft))X»Y12.IdG] (T+0exp(-ft)) 143 144 fijdGGKT-Wexpértnxly, z I 6(T+eexp(-T‘t)) expEJrIe G'EGXJ Béexpi FEW (T+°°XP(‘”))]de’ 6] (2-A) rT+9exp(~l‘t) e 11 and 12 of equation (4n’7) may be written in spherical polay coordinates as I " ' 1 = I F (w,F,T) c0520 sinw dw TI 2 =I F (12.1.31) 511130 dd» 0 where F(w’f"T‘T§J‘['\,—;Lf_‘g+ 1+2a ( 3‘3 Afcos I! + A ,,_s.1.nzg+III)/2 (A,cos 111 + Alsin I») (Agcos’w+ Aisin’tp )'/Z 2a2 +(A’;cos up + Xzsin ’wfi A;COSZIIJ+ Aisin 0) l/2a2 - (A’tosz I» + A 2sin writ/K). sinzw 4» $3111 11»)ij (5 A) 2 Y = eXP-zafl COS’W + 12.91111 1/2 W] A1 5 [e-ZFT + (10.2%)]15 L I. 21".- 42 = [e T 1‘? (e11T - 013 1‘ A3 : e-1‘T, A“ : EFT/2 T = Bot, T, = 80(t-t’) and f = F/Bo 145 By substituting w = tan 0 in the above equations, we obtain ,’ 1~ :‘I =[;W(2~w1 2 I . 2 ;q31%r_jfw . 0 V + 2w 2 {23+ Alw) “3+ 3E0”) + (01+ i” A3+ kw)” - U452 exp-2a 11 + 1‘ w 1/2 (7-A) 01+ 111020? 801 [8732?] Then letting _ 1+(X /A )Zw Vz ’ Hui/13W ’ (M) we obtain with integration by parts, Ifis - I = l -2a/ = det(l)[(” dodndzg(o,n,Z)fo( eXp(-17mt) nexp(l—-"21t,z) —oo -2T A [:(T + 0exp (-J§1t) )510121de 0(T+0exp(-% t)) oo 00 + -m' - - + % [I] dpdndzg(p,n,z) d0G((T 0exp( 13101012) -w 1 0(T+0exp(-lfllt)) 0 im A _ (T+0exp(--—-t)) , '-J 8(9 D, 0 2 2 9 Z: N) 99-61] (8'2) 1+eexp(—‘—'g— t) All moments , , , , and are gen- erated by eq.(B-2). Since in uniaxial shear flows no deformation occurs in z-direction, z is arbitrarily set to zero, then these moment integrals are evaluated as exemplified by integral <02> = d6t(T) ‘89-“ 03%) 9.1: J! dpdnpzexplg 3_2N(}\2102_2wpnq+>\:n2) 0 -oo 2a2/3N AipZ-Zan+A:vz 149 T _ I I 2 J dt'e T J[ dodnEXP[3%¥(A1 02-2W4'00+A:nz) o -oo 2 2a /3N :] (B_3) A5202 - 2Won + X1202 Where = -imt .3; _ -mt % 11 (e + im (1 e )) . ' 1 2 A2 _ (EImT + % (eImT_-I))2, A“ : eImT (B_4) It is noted here that A's in 2nd term of eq(B-2) are defined as in eq.(B-4), however, they are functions of elapsed time, T'. Next the p(p,n) frame is transformed into cylindrical polar coordinates h(d,0). On intergrating out the radical component d, eq.. (B-4) becomes = det(I)(§—)(§;) { e"T I I (1,0) sinzwdw T,-T' .2 I _ + dT e I(A,w)51n vdWI (B 5) 150 where K2(U) 2 (32_ 10,41) = 9N2 (AisinZw-Zquinwcosw+A§c052¢)(A:sin20—2 sinwcosw+AEcoszw) K2(u) is a 2nd order Bessel function, and 2a(A§sinw—Zquinwcosw+A%coszw )% A§sinw - 2Wsinwcosw + Ajcoszw Since a <0(1), p is small from the A expressions. Then the series for the Bessel function of integral order and 0f the 3rd kind is util— ized in order to completely integrate out the coordinate variable,i.e., K ( ) = 1n§1(_1)k 10:5:11L___ + ( 1)n+1% (U/2)n+2kLInU/2-%W(k+1) n U 2 = 1 n-2k — = I I K o k.(p/2) k 0 k.(n+k). -%v(n+k+1)] (B-6) where v(.) is the Euler's psi function. Thus applying eq. (B—6) to eq.. (5—1) we obtain Zn I(0,T) =-% Jd¢SIPV ( O l Aisin20-2quinwcos¢+X:coszw)2 a2 (Aisinzw-quinwcosw + Aficoszw)(Afisinzw-ZWsinwcosw + AEcosE) + 0(a4) (B—7) 151 The first term in Eq (B-7) is analytically integrable, but the second term is integrated by the method of partial fraction to yield a final expression of 1(1) as A; 2a2(811§ - A1Wq) (B-8) 1(1)=-——1—~-————r—.— (1212 - quZI/Z xi (111% - wqu) / 12 where A1 = ~2WIAiAE- Afiq)/C and c = 4112003 - (10315 + A012.) + 12) — (2103 - 0311+ 1111)) Thus the moment integral is obtained as APPENDIX C OSCILLATORY SHEAR FLOWS—-TRANSFORMATIONS AND CALCULATIONS 152 APPENDIX C OSCILLATORY SHEAR FLOWS-~TRANSFORMATIONS AND CALCULATIONS The transformation to the sinusoidal domain with u = sinmt after the characteristics of eq. (5.21) have been defined gives a slightly different o.d.e. of the form A df + Bf _ G (H) dU (1-U2)% (1-U2)% The solution of this by method of characteristics is u du'B 1m u' -m u' f(psnazsu) = f0(po’wo’zo)eXp_ w(1_ul2 % (DOeXP(jf——),UEXP( 2 ),20 o - -1 Sln u A .~ . .c . du'G imou. -1mou + ———_ (0 eXp( )m exp(———),z ) J (1_u|2)éw O 2 O 2 0 o u A imp" -m0uK\ x exp - du"B(pOeXp(—25——), noexp( -§5——), 20) (C-2) l u Using transformations identical to Appendix B, except the independent variable is u' instead of t'. The moment integral is obtained as 154 ~ ~ ~ w -Imou -im u <9(0,n,Z)> = d8t(T) dodndzg(o,n,Z)fo [peXp( w ).neXP( : )’ZI 1' 00A III] U _ _ T B[(T+0exp(- g )0.n,Z)d9 x exp - . ~ . (Imow -im0u 2 wln (T+eexp(lmgfl) 2 2 J -' ——‘ ———__—— 1 0(T+0exp( w )) 1+u imo ( 0 ) w m A i~ou)_ _ d06((T+0exp(- 0.0.2) + l%— dodnd29(p,n.Z) w . 1m M00 0 _m 1 imou l+u2_g;g ((T+ °Xp( w 0(T+0expC-—73-)) imo 0 e ( > T+0ex — o .1 A ,- ( ———7—J3 ),z,N)de' x exp 10;; B(6 0.0 9 .~ (C-3) .- . 12 . 'mo“ 2 i imou 0 (1+ 57— (ln01/e- ) ) T+eexp(- 0 w In oscillatory shear flows, the transients areallowed to die out consequently the first term in the moment expressions damps out. Pertinent moments can be derived from eq. (C-3) from the specific of choice G and B. The moment <02> will serve as an example. 155 2 = EL .EN _ - <0 > dEt(l) 80 (Zn) [di'e T dpdnp2 o -oo 2 exp _ 3—2N(p2>\'2 _‘ qu'pn‘l’Aznz)" ‘_.l§-_/3_N_'_. ((3-4) 1 2 A202_2wpn+)\:n2 ' 3 where u im 2 1m e-L—J2 (u—u")]du" A1 — exp - ——9 (U'U') + 25m w 1 (0 0 (1_ulI2)2 uI u'im _0 _ II 1110 J1. (””11 du" = ———— - ' + ———7— A EXP (u u ) 21m (1-u"2)% u 2 IITIO (U‘U') 2 IIIIO A3 — exp-TI— X1 — exp ~Zr-(u-u ) q'= 1+8T’ m = m /B t' = 80(t-t') Utilizing transformation procedures and integration techniques identical to those used to obtain eq. (B—9), we obtain the ocillatory shear function as i: " E I kidet(f)(%1—~—) e'T 1(1) 111! (C-5) ~ -5 e23 1+2a Sxy = é; O 156 where A2 - A2 1(1) . 1 2 _ 5133 01+ 1%)0113- 1.213) I 2 2 3/2 8 (AiA ' qzwz) .2 2 2 2 2 2 - 21 [(A1 - 1,)11 + w (A. - 13)] (C-6) C = 41420030030 + mi + C12) - M2013 - (1315 + 11111) In the first term xi _ A3 = -2i[5in m0(sin11t - sin(t— t'/BO)) u E SInIfiO