1. Trang chủ
  2. » Kỹ Thuật - Công Nghệ

Thermosetting Polymers Part 14 potx

48 121 0

Đang tải... (xem toàn văn)

Tài liệu hạn chế xem trước, để xem đầy đủ mời bạn chọn Tải xuống

THÔNG TIN TÀI LIỆU

Thông tin cơ bản

Định dạng
Số trang 48
Dung lượng 352,19 KB

Các công cụ chuyển đổi và chỉnh sửa cho tài liệu này

Nội dung

The behavior of the material can be characterized bytwo quantities: the equilibrium water concentration W1, which charac-terizes the polymer affinity for water hydrophilicity, and the dur

Trang 1

Durability

Aging can be defined as a slow and irreversible (in use conditions) variation

of a material’s structure, morphology and/or composition, leading to adeleterious change of use properties The cause of this change can be theproper material’s instability in use conditions or its interaction with theenvironment (oxidation, hydrolysis, photochemical, radiochemical, or bio-chemical reactions, etc.)

Aging becomes a difficult problem to study in practice, because itproceeds too slow in use conditions (typical lifetime of years) It is thennecessary to make accelerated aging tests to build kinetic models thatdescribe the time changes of the material’s behaviour, and to use thesemodels to predict the durability from a conventional lifetime criterion.Indeed, the pertinence of the choice of accelerated aging conditions, themathematical form of kinetic model, and lifetime criterion has to be proved.Empirical models are highly questionable in this domain because they have

to be used in extrapolations for which they are not appropriate

In the ideal case, an aging study would involve the following steps:(a) Identification of aging mechanisms through physical and analy-tical observations

Trang 2

(b) Kinetic modeling based on the previously established mechanisticscheme (diffusion law, chemical kinetics, etc.).

(c) Prediction of use properties from the structural state using mer physics

poly-Among polymers, thermosets are especially difficult to study for manyreasons: structural complexity, making difficult the chemical analysis, lack

of rigorous tools to investigate the macromolecular structure; lack of sical theories to interpret the change of properties (e.g., embrittlement)against structural changes

phy-These difficulties are again increased, sometimes considerably, by thefact that thermosets are generally used in composites or as adhesives, e.g., inapplications where aging can result also from a change of the interfacialproperties and in which certain key properties, e.g., the rubbery modulus,are practically inaccessible

Despite these difficulties, the practical importance of durability incomposite or adhesive applications, has given rise to a vast amount ofliterature in this field during the past decades Most of the studies dealwith two main aging cases:

1 ‘‘humid aging’’ in liquid media (boats, pipes, tanks, etc.) or in wetatmospheres (aerospace structural parts, (e.g., helicopter blades)

2 ‘‘thermal aging’’ at high temperatures in processing as well as inuse conditions (engine parts, electrical insulations, etc.)

Both domains constitute the main sections of this chapter

Photochemical or radiochemical aging are not examined here formany reasons:

1 The literature on thermosets in these domains is very scarce.Most authors use theoretical concepts and experimental methodselaborated for linear polymers, and extensive reviews are avail-able (Ranby and Rabek, 1975)

2 For photochemical aging, it is well known that photooxidationaffects only a thin superficial layer directly exposed to solarradiation – a few dozens of micrometers in the case of epoxies(Bellenger and Verdu, 1983) Thus the aging mode cannot con-trol the material’s lifetime in most cases (composites, adhesives),except for applications such as, for example, varnishes of auto-motive bodies (Bauer et al., 1992)

3 Radiochemical aging has very specific applications The uses ofthermosets in nuclear engineering have been growing Most ofthe data on radiochemical aging of thermosets, available at thebeginning of the 1990s, has been reviewed (Wilski, 1991)

Trang 3

14.2 AGING RESULTING FROM WATER

expos-In case (a), an equilibrium is reached It can be considered here thatthere are only reversible physical interactions between the polymer andwater Drying leads to a curve that is practically a mirror image of theabsorption curve The behavior of the material can be characterized bytwo quantities: the equilibrium water concentration W1, which charac-terizes the polymer affinity for water (hydrophilicity), and the duration tD

of the transient, which is sharply linked to the sample thickness L and to aparameter characteristic of the rate of transport of water molecules in thepolymer – the diffusion coefficient D

In cases (b) and (b0) there is no equilibrium; the mass increases tinuously or decreases after a maximum, which indicates the existence of anirreversible process – chemical, hydrolysis, or physical, damage; microcavi-tation increases the capacity of the material for water sorption The experi-mental curves having the shape of curves (b) or (b0), indicate that theirreversible processes induce significant mass changes in the timescale ofdiffusion When the irreversible processes are significantly slower than dif-fusion, the behavior shown by curves (c) or (d) is observed The sorptionequilibrium is reached at t  tD, and a plateau can be observed in the curve

con-FIGURE14.1 Shape of the kinetic curves of mass change in the most frequentcases of humid aging

Trang 4

before irreversible changes become significant.

Since both reversible and irreversible processes are influenced in tinct ways by temperature and water activity, the first step of a humid agingstudy consists of searching for the conditions (T, RH, sample thickness) inwhich both phenomena can be clearly decoupled, as inFigs 14.1candd.Theinterpretation of experimental results and the modeling of the kinetics ofproperty changes would be difficult or even impossible if physical character-istics such as W1and tD (or better D) were not known

dis-To distinguish between physical and chemical aging, one needs tical data on structural changes (chain scission by hydrolysis, decrease ofcrosslink density, evolution of small molecules resulting from degradation).Visual and microscopic observations enable us to detect damage In themost complicated case, chemical degradation and mechanical damage aresharply coupled (osmotic cracking) The following section is devoted tophysical, reversible, water–polymer interactions and water diffusion

analy-14.2.2 Hydrophilicity

Polymer hydrophilicity can be judged in the cases (a), (c), or (d) of Fig 14 1

It is defined as the affinity of a polymer for water, which can be quantified

by the equilibrium mass gain, W1, determined in standard conditions, e.g.,

in a saturated atmosphere from a sorption experiment W1depends on thevapor pressure or activity of water and on the temperature It varies, typi-cally from 0 to 10% in most networks

a Influence of Vapor Pressure or Activity of Water

W1 is usually obtained by weighing:

W1¼Asymptotic mass  initial mass

The water equilibrium mass fraction is given by

m1¼ W1

For rough estimations, one can use m1 W1

The water equilibrium concentration is expressed by

Trang 5

mechanism For a low-to-moderate hydrophilic behavior, it may be assumedthat Henry’s law is valid, at least in a first approximation:

where S is the solubility coefficient

Thus, in the domain of validity of Henry’s law, the equilibrium centration is proportional to the relative hygrometry (at a given tempera-ture)

con-Immersion in pure water must lead to the same result as in a saturatedatmosphere If water contains solutes, its vapor pressure decreases Thecorresponding equilibrium concentration, linked to the water activity, isproportional to the water vapor pressure In other words, the water equili-brium concentration is a decreasing function of the solute concentration:salt water is less active than pure water

Most of these aspects of water-sorption equilibrium correspond to theequality of chemical potentials of water in the medium and in the polymer.The consequences of this principle are illustrated by the experiment of Fig.14.2, where an interface is created between water and a nonmiscible liquid(oil, hydrocarbon, etc.), and a polymer sample is immersed into the organicliquid It can be observed that, despite the hydrophobic character of thesurrounding medium, the sample reaches the same level of water saturation

as in direct water immersion or in a saturated atmosphere What controlsthe water concentration in the polymer is the ratio C=Cs of water concen-trations in the organic phase, where Cs is the equilibrium concentration,which can be very low but not zero In other words, hydrophobic surfacetreatments can delay the time to reach sorption equilibrium but they cannotavoid the water absorption by the substrate

Let us remind ourselves that the saturated vapor pressure, ps, increaseswith temperature In a first approximation it can be written:

FIGURE 14.2 Water equilibrium in a nonaqueous medium immiscible withwater

Trang 6

p ¼ p0exp Hw

where ln p0¼25:33 (p0 in pa) and Hw¼42:8 kJ mol1

On the other hand, S obeys an Arrhenius law:

S ¼ Sexp Hs

where Hs is the heat of dissolution In the case of water, Hsis negative (thedissolution process is exothermic), and ranges generally from 25 kJ mol1(polymers of low polarity) to 50 kJ mol1 (highly polar polymers) Then,

C1¼Sp ¼ S0p0exp ðHsþHwÞ

For highly polar polymers, Hs< Hw, and the equilibrium tration is a decreasing function of temperature This is often found in themost hydrophilic networks, based, for example, on the aromatic amine –aliphatic diepoxide of diglycidyl ether of butane diol (DGEBD) type(Tcharkhtchi et al., 2000), or on particular polyimides (Hilaire and Verdu,1993)

concen-For many usual, moderately polar networks, such as epoxides of thediglycidyl ether of bisphenol A (DGEBA) diamine type, or vinyl esters,

Hs Hw, so that the equilibrium concentration appears almost ture-independent For most of the less polar networks such as polyesters oranhydride-cured epoxies, C1 (or W1) increases slightly with temperature:

tempera-W1=T 0:01–0:02% K1 between 20 and 50 C

Except, eventually, for networks of very low polarity, Hsis close to

Hw, so that the above equation can be well approximated by

Trang 7

tions, so that there is no reason to assume that water is not homogeneouslydissolved into the polymer matrix, except, in the case of a macroscopicporosity.

From a global analysis of W1 values (Table 14.1), it is clear thathydrophilicity is an increasing function of the polarity of the groups con-tained in the polymer and their concentration:

W1¼1:63 Kð u4:5ÞðW1in %, Ku in GPaÞ ð14:9ÞThe intercept (4.5 GPa) corresponds essentially to the nonhydrophilicdispersive component of cohesive energy, which does not vary very muchfrom one polymer to another

Predictions of W1 could be performed using global (Hildebrand) orpartial (Hansen) solubility parameters but these are very difficult (andperhaps impossible) to determine accurately from solvent–sorption experi-ments, so that this way is not realistic The best experimental approach is, inour opinion, using the ultrasonic modulus

Empirical calculations of W1 from the CRU structure can give tively good results provided that sufficiently large structural units are con-sidered to take into account eventual intramolecular interactions (Bellenger

rela-et al., 1988) However, their practical interest for predicting the behavior ofnew systems is relatively limited: the molar contribution of the most hydro-philic groups, e.g., hydroxyl groups, is not an integer, which means that

Trang 8

TABLE14.1 Tentative classification of network hydrophilicities

Very low hydrophilicity Polydimethylsiloxane Typically < 0:5%

PolyethylenePolystyrene co divinylbenzene

Low hydrophilicity Many styrene crosslinked UP 0.5–1.5%

Some styrene crosslinkedvinylesters

Increase with esterconcentrationSome anhydride crosslinked

epoxiesModerate hydrophilicity Some vinyl esters 1.5–3%

Amine crosslinked epoxies ofrelatively low crosslinkdensity

Increases withalcohol or amideconcentrationSome polyimides

High hydrophilicity Amine crosslinked epoxies of

high crosslink density (TGAP,TGMDA)

> 3%

Many polyimides

FIGURE14.3 Correlation K–W1between the bulk modulus and the equilibriumwater concentration o Amine crosslinked epoxies Vinylesters and poly-esters

Trang 9

certain groups are able to fixate a water molecule and others are not Atheory for predicting the fraction of ‘‘active’’ groups is necessary.

Recent studies showed that, in a given structural series in which themain variable is the OH concentration, the molar contribution of a parti-cular OH group is apparently an increasing function of the OH concentra-tion A possible explanation of this result is that in the polymer–watercomplex, water is generally doubly or triply bonded Thus, a hydrophilicsite would be composed of two, eventually three, neighboring polar groups.The average distance between polar groups is generally too high to allowsuch a concerted process But the spatial distribution of these groups is more

or less heterogeneous, so that there is a more or less important fraction ofthese groups sufficiently close to form an hydrophilic site Thus it is possible

to explain why OH groups have a weak molar contribution to water tion in polyesters (½OH  0:5 mol kg1) (Bellenger et al., 1990), a mediummolar contribution in amine-crosslinked epoxies (4 < ½OH < 8 mol kg1)(Bellenger et al., 1988); and a high molar contribution in linear polymerssuch as those soluble in water – poly(vinyl alcohol), poly(acrylic acid), etc –

absorp-in which ½OH > 10 mol kg1 It explains also why, in epoxide–amine tural series differing by the amine/epoxide molar ratio, W1 increasespseudo-parabolically with the OH concentration (Tcharkhtchi et al., 2000)

struc-14.2.3 Diffusion

Solvent transport in organic polymer matrices is usually depicted as a step mechanism The first step is the dissolution of the solvent in the super-ficial polymer layer This process, which can be considered almost instanta-neous in the case of water, creates a concentration gradient The second step

two-is the diffusion of the solvent in the direction of the concentration gradient.This process may be described by a differential mass balance (often calledFick’s second law), which, in the unidimensional case, may be written as

1ð2n þ 1Þ2exp 

Trang 10

D ¼

16L

2ðC=C1Þ

ð ffiffit

p

A characteristic time of diffusion, tD, defined as the duration of thetransient, can be arbitrarily taken at the intersection between the tangent atthe origin and the asymptote(Fig 14.4).This leads to

Typical values of diffusion coefficients are 1012–1013m2s1 at 20–

50 C The diffusion time tDis about 1 day to 1 week for samples of 1 mmthickness, and 1 year for samples of 1 cm thickness

Here, the best way to accelerate aging is to decrease the sample ness (when possible) Except in very highly hydrophilic materials, D is inde-pendent of the relative hygrometry or water activity, but is an increasingfunction of temperature:

Trang 11

The structure–diffusivity relationships are not clearly established Arough correlation between the diffusivity and the reciprocal of hydrophili-city is usually found For example, for networks absorbing less than 1.5%water, such as unsaturated polyesters and anhydride-cured epoxies, the

13

m2s1 at 20–50 C

In contrast, for epoxies cured by amines absorbing more than 2% water,the coefficient of diffusion is of the order of 1013m2s1at 20–50 C, and ittends to decrease with W1 in most of the structural series

Diffusion kinetics were studied on both sides of Tg for epoxy samplesbased on an aliphatic diepoxide (DGEBD) cured by an aromatic diamine(diethylenetoluenediamine: ETHA) Surprisingly, there is no clear disconti-nuity of D and S around Tg (Tcharkhtchi et al., 2000) All these observa-tions suggest that, in these systems, the diffusion rate of water is controlled

by the strength of the polymer–water hydrogen bond, as in the followingscheme:

hydro-14.2.4 Physical Consequences of Water Absorption

a Plasticization

Using the classical hypotheses of the free volume theory, the glass transitiontemperature for a polymer (p) and solvent (s) solution, with a volume frac-tion of solvent, v, is given by (Kelley and Bueche, 1960):

Tg¼ð1  vÞpTgpþvsTgs

where Tgp and Tgs are the glass transition temperatures of the dry polymerand the solvent (Tgs100–150 K for water, depending on authors), and pand sare the respective coefficients of free volume expansion ð ¼ 1gÞ.This relationship can be simplified assuming the validity of the Simha–Boyer rule, T ¼constant, which leads to

Trang 12

With the equilibrium water content It is almost insignificant in polarity networks such as unsaturated polyesters or anhydride-cured epoxies.

low- With the initial glass transition temperature Tgp of the polymer inthe dry state It is limited in low-Tg epoxies and vinyl esters, butimportant in high-Tg amine-cured epoxies

b Swelling

The swelling mechanism is well established in rubbers (Flory, 1943) Theexpansion force generated by the solvent penetration is equilibrated by theentropic force linked to chain stretching Little is known, in contrast, for thecase of glassy polymers where plasticization effects are not sufficient toinduce a devitrification Swelling can be defined by

V

where V0 and V are the sample volumes in the dry and wet (equilibrium)states, respectively; is generally lower than the value predicted from thehypothesis of additivity of polymer and water volumes This means thatwater penetration in the polymer is accompanied by some contraction

To our knowledge there is no theory that predicts efficiently the ling ratio from W1 and the network characteristics

swel-TABLE14.2 Calculated plasticization effect in saturation conditions for someindustrial networks

Trang 13

14.2.5 Mechanical Consequences of Water

Absorption

a Plasticization Effects

In the domain where the material is ductile or semiductile, the yield stress

yis linked to the glass transition temperature Tg by

y ¼CkðTgTÞ

where Ckis a constant (Kambour, 1984;Chapter 12).Plasticization, whichdecreases Tg, leads to a decrease of y This effect can be important when amaterial that is not very far from its glass transition point is submitted toloads In this case, water absorption can induce yielding The data ofTable14.2show that ycan decrease by more than 30% A significant increase ofcreep compliance can be expected in such a case

A softening can be observed at temperatures ranging typically from Tg(dry) to (Tg (wet) – 50 K) Far from Tg there is no significant influence ofwater (at moderate concentrations) on stiffness

equi-(a) Establish the diffusion law From this law, calculate the waterconcentration C at any time and any point of the sample.(b) From equilibrium experiments, determine the relationshipbetween the concentration C and the degree of swelling V=V.(c) From the data obtained in (a) and (b), calculate the volumechange at any point

(d) From the profile of volume change, calculate the strain profile

Trang 14

(e) From the behavior law  ¼ fð ; t; CÞ, calculate the stress Indeed,the effect of sorbed water on the behavior law has to be estab-lished under equilibrium conditions.

The results in terms of stress distribution along the thickness areshown in Fig 14.5

In extreme conditions, e.g., essentially at temperatures typically higherthan 60 C, where diffusion is fast and generates strong concentration gra-dients and where the yield stress is sufficiently low, water absorption caninduce damage in medium to high hydrophilic networks

14.3.1 The‘‘Ideal’’ Caseof Hydrolysis

Hydrolytic processes can be analyzed by comparison with an ideal case,which can be defined by the following hypotheses:

FIGURE 14.5 Stress distribution at various stages of water absorption anddesorption

Trang 15

(a) Hydrolysis is homogeneous at all dimension scales It is not sion-controlled (no degradation gradient along the thickness).

diffu-(b) Hydrolysis (E þ W ! chain scission) obeys a second-order kineticlaw:

dn

where n is the number of hydrolysis events (moles) per unit volume, E is theconcentration of reactive groups, C is the water concentration, and k is therate constant that depends only on temperature C is an increasing function

of the hygrometric ratio in a first approximation:

con-(d) A pertinent end-life criterion may be defined, corresponding to aparticular conversion (e.g., to a given structural state independent of expo-sure conditions): n ¼ nf Then,

Trang 16

The lifetime may be thus calculated as

264

37

14.3.2 Determination of The Conversion Degree in

Thermosets

Infrared or NMR Determination of ½E These methods have beenused for instance in unsaturated polyesters or in vinyl esters (Ganem etal., 1994); however, they are not very sensitive at low conversions

Rubber Elasticity This method is based on the hypothesis thatFlory’s relationship can be applied to thermosets, which is only a roughapproximation(Chapter 10):

where G is the shear modulus at temperature T (above Tg e is theconcentration of elastically active network chains For a network initiallyfree of dangling chains, each chain scission leads to the disappearance of oneelastically active chain, for networks in which the crosslink functionality is

’ 4, and three elastically active chains for networks in which the crosslinkfunctionality is ’ ¼ 3(Fig 14.6).Then,

Trang 17

e¼ 1

3RTðG0GÞ for ’ ¼ 3 ð14:29bÞ

In the case of unsaturated polyesters, nondegraded samples made from

a prepolymer of molar mass M and a styrene mass fraction s have a

chain-e is theactual concentration of elastically active network chains, an ideal networkwould be obtained by ‘‘welding’’ each chain end to another one, leading to

FIGURE14.6 Destruction of elastically active network chains resulting from achain scission in the case of tetrafunctional (a) and trifunctional nodes (b)

Trang 18

This approach can be illustrated by unsaturated polyesters based on

an almost equimolar combination of maleate and phthalate of propyleneglycol, crosslinked by styrene (45 wt%) (Mortaigne et al., 1992) Six samplesdiffering by the prepolymer molar mass were analyzed The chain-endsconcentration, b, was determined by volumetric analysis of alcohols andacids in the initial reactive mixture Then, the system was cured, elasticmeasurements were made in the rubbery state at Tgþ30 C, and the shearmodulus G0 was plotted against chain-ends concentration (Fig 14.7).Thefollowing relationship was obtained:

Trang 19

styr-For example, unsaturated polyesters based on neopentyl glycol aremore stable than those based on propylene glycol In series differing only

by the diol nature, a good correlation between the hydrophilicity and thehydrolytic stability is generally observed, as illustrated by the example ofFig 14.8 The pseudo-first-order rate constant of hydrolysis (K ¼ k½E0Þ,was plotted against equilibrium water concentration for two linear polyesterfamilies: the first one based on isophthalic acid (IPA) and the second onebased on maleic anhydride (MAA) For a given family, the hydrophilicitydepends essentially on the diol structure, the order being always: neopentylglycol ðNPGÞ < propylene glycol ðPGÞ < diethylene glycol (DEG)

The intrinsic reactivity of the ester, which essentially influences k,plays an important role The following order of increasing stability hasbeen established: reacted fumarate < orthophthalate < isophthalate methacrylate It can be seen in Fig 14.8 that maleates are about 10 timesmore reactive than isophthalates (Belan 1997)

Practitioners know well that isophthalates are more stable thanorthophthalates, and that vinyl esters (methacrylates) are at least oneorder of magnitude more stable than unsaturated polyesters (UP) Theweak point of UP is the fumarate unit, which is, however, necessary forthe crosslinking

FIGURE14.8 Pseudo-first-order hydrolysis rate constant at 70 C against librium water concentration for isophthalate (*) and maleate (~) polyesters.The abbreviations are given in the text

Trang 20

equi-An estimation of ester concentrations and rate constants for somepolyesters and vinyl esters crosslinked by styrene is given in Table 14.3.

To compare with linear polymers: rđ100 5mol kg1 day1, so

5

day1 for bisphenol A polycarbonate đơE04 mol

kg1), and rđ100 3 mol kg1 day1

103 day1 for poly(ethylene terephthalate) đơE05:20 mol kg1)

It appears that the ester reactivity towards hydrolysis is governed bythe chemical environment of the ester group rather than by the macromo-lecular architecture (there is no systematic difference between linear poly-mers and networks)

In polyesters, the existence of an osmotic cracking process (see below),gives importance to other structural factors, especially the concentration ofdangling chain ends

TABLE14.3 Characteristics of the hydrolysis kinetics at 100 C, 100% HR forstyrene crosslinked unsaturated polyester and vinyl esters (After Ganem

Trang 21

thermosets are initially brittle In this case, linear elastic fracture mechanicscan be applied The ultimate stress, f, may be related to the modulus E by

polye-Thus, there are two ways to select an end-life criterium for thermosets: arbitrary choice in the case where there is no cracking

cracks appearance in the other cases

It remains then to establish the relationship between the hydrolysisconversion and the crack initiation, which is not obvious

14.3.5 Some ‘‘Nonideal’’ Cases of Hydrolytic Aging

a Diffusion-Controlled Hydrolysis

The kinetic equations analyzed in the previous sections are valid for aspatially homogeneous hydrolysis However, it is possible that above a cer-tain boundary in the reactivity (K) and sample thickness (L) map, all thewater molecules penetrating in the sample are consumed in superficiallayers, so that aging becomes heterogeneous(Fig 14.9)

A simplified theory of these processes has been established (Audouin

et al., 1994) It leads to a simple scaling law for the prediction of the ness of the degraded layer (TDL):

thick-TDL

ffiffiffiffiffiffiD

There is a limiting thickness, Lc2TDL, such that for L < Lc, aging

is homogeneous, whereas for L > Lc, it is heterogeneous (the degradationrate is lower in the core than in the superficial layers) From

Trang 22

HD is of the order of 20–70 kJ mol1; HR is generally higher (50–

100 kJ mol1), so that HL is generally negative This means that the ness of the degraded layer is generally a decreasing function of temperature,

thick-as is experimentally observed in many cthick-ases A practical consequence of thisresult is that care must be taken in the choice of the combination of samplethickness and exposure conditions, for accelerated aging

b Nonzero-Order Processes

Sometimes, the experimental n ¼ fðtÞ curve is not linear in its initial part: forexample, in the case of polycarbonate, where the rate of chain scissiondecreases rapidly in the early period of exposure and tends towards a con-stant value, r1, so that: n ¼ n1þr1t In such cases, there is no other way toexplain this behavior than to assume that the material contains veryunstable groups (in a concentration close to n ¼ n1) that degrade rapidly.When they are consumed, the system adopts a ‘‘normal’’ pseudo-zero-orderbehavior

FIGURE14.9 Non diffusion controlled and diffusion controlled domains in areactivity–thickness (K–L) graph The thickness profiles of degradation at var-ious points on both sides and at various distances of the boundary B areshown

Trang 23

c Random and Nonrandom Hydrolysis – Mass Changes

Gravimetric studies of hydrolysis are especially interesting because theycombine simplicity and richness of information provided that a kineticmodel is available to interpret experimental curves This model can bebuilt from the following approach: let us consider a network containinginitially ½E0reactive linkages, bodangling chains (each one containing initi-ally  reactive groups sufficiently near to the chain end to give a small,extractable molecule by hydrolysis), and 0 free extractable molecules(monomers), such as unreacted glycols, etc The rate of scission of elasticallyactive chains (or dangling chains far from their extremity) is

dn

dt ¼K ½E0

ð14:41Þ

where b is the actual concentration of dangling chains These processes lead

to two new chain ends, so that

where Mnb is the number-average molar mass of extractable molecules in

g mol1 The whole mass change is calculated from

Trang 24

In the case of a true random process, all the groups are equireactive, sothat K ¼ K0, and

18½E0

M  0

ð14:55Þ

... calculate the strain profile

Trang 14< /span>

(e) From the behavior law  ẳ f ; t; Cị, calculate the stress Indeed,the...

14. 3.1 The‘‘Ideal’’ Caseof Hydrolysis

Hydrolytic processes can be analyzed by comparison with an ideal case,which can be defined by the following hypotheses:

FIGURE 14. 5... constants for somepolyesters and vinyl esters crosslinked by styrene is given in Table 14. 3.

To compare with linear polymers: rđ100 5mol kg1 day1,

Ngày đăng: 11/08/2014, 06:21

TỪ KHÓA LIÊN QUAN