• Aucun résultat trouvé

Critical behavior of trifunctional randomly branched polycyanurates

N/A
N/A
Protected

Academic year: 2021

Partager "Critical behavior of trifunctional randomly branched polycyanurates"

Copied!
12
0
0

Texte intégral

(1)

HAL Id: jpa-00247691

https://hal.archives-ouvertes.fr/jpa-00247691

Submitted on 1 Jan 1992

HAL is a multi-disciplinary open access archive for the deposit and dissemination of sci- entific research documents, whether they are pub- lished or not. The documents may come from teaching and research institutions in France or abroad, or from public or private research centers.

L’archive ouverte pluridisciplinaire HAL, est destinée au dépôt et à la diffusion de documents scientifiques de niveau recherche, publiés ou non, émanant des établissements d’enseignement et de recherche français ou étrangers, des laboratoires publics ou privés.

Critical behavior of trifunctional randomly branched polycyanurates

Jörg Bauer, Walther Burchard

To cite this version:

Jörg Bauer, Walther Burchard. Critical behavior of trifunctional randomly branched polycyanu- rates. Journal de Physique II, EDP Sciences, 1992, 2 (5), pp.1053-1063. �10.1051/jp2:1992185�.

�jpa-00247691�

(2)

Classification Physics Abstracts

82.35 05.50

Critical behavior of trifunctional randomly branched

polycyanurates

J6rg Bauer (*) and Walther Burchard

Institute of Macromolecular Chemistry, University of Freiburg, W-7800 Freiburg, Germany (Received 9 January 1992, accepted 31January 1992)

Abstract.-The behavior near the gelation threshold of trifunctional randomly branched

polycyanurates is studied by static and dynamic light scattering. By static measurements the critical exponents y, a and v were obtained, which describe the divergence of the weight average (M~) and the characteristic (M* ) molecular weights and the radius of gyration (R~) respectively.

All these independently measured exponents together with T, describing the power law behavior of the molecular weight distribution and measured by size exclusion chromatography coupled

with light scattering, confirm the predictions of the three-dimensional percolation theory.

Furthermore, it was shown by dynamic light scattering that the power law behavior over some decades in time of the time autocorrelation function and the divergence of the mean relaxation time are characteristics of the gelpoint. The development with increasing reaction time of the time

correlation function of the gelling system from the pregel through the gel-point into the gel state was analyzed quantitatively by a hybride of a stretched exponential and a power law function. It

was shown that the basic relations between y, ~, v and T are fulfilled and should be universally

valid. On the other side, the relations, which link the characteristic exponents of the time correlation function at the gelpoint to the basic exponents, gave incoherent predictions. In that

way, one can assume that their universal validity is limited, since some additional assumptions on

the local dynamics are needed.

Introduction.

In a previous paper [Ii we introduced polycyanurates, the polycyclotrimerization products of difunctional aromatic cyanic acid esters, as a model for trifunctional, randomly branched

polymers [2, 3]. As a first result the scaling relations

R~ M$ R~ M$ with p'

=

0.52 ± 0.03 (1)

for the radius of gyration R~ m ((s~) )~'~ and the hydrodynamic radius R~

m ( I/R )

~

)~ were

(*) Permanent address Fraunhofer Institute of Applied Materials Science, O-1530 Teltow-Seehof, Gernlany.

(3)

obtained by static and dynamic light scattering. Furthermore, the predicted power law behavior of the weight fraction molecular weight distribution [4]

w (M ) M ~ f (M/M *) with T

=

2.22 ± 0.02 (2)

was found using size exclusion chromatography coupled with a low angle light scattering detector, where f is a cutoff function and M* a characteristic cutoff molecular weight. The exponent p' contains informations on both the molecular weight distribution and the solvent

quality and is related to p for the molecular uniform clusters by p = p'(3 T), which gives

p = 0.4I and a fractal dimension D

=

I/p

=

2.44. The excellent agreement of T with the

percolation prediction of 2.2 suggests that three-dimensional percolation is a better theory for

the critical region than the classical Flory-Stockmayer approach [5, 6] from which

T = 2.5 results.

On the other side, we found that the classical theory in its extension by Gordon's cascade

theory [7-9] works well for predicting the scattering behavior of the polycyanurate solutions.

Although the assumption of Gaussian chain statistics, which ignores all volume effects, seems

to be unrealistic, the calculated scattering curves are in excellent agreement with experimental

data. Apparently the fractal dimension D

= 4.0 of a single cluster is counter-balanced, if we average over the broader molecular weight distribution resulting from the Flory-Stockmayer approach.

Thus, it appears possible to use the polycyanurates with their simple, selective chemistry and their well defined broad molecular weight distribution for checking the predictions of both the classical and the three-dimensional percolation theory. The present paper now deals with the critical behavior of this polymeric system in two ways : by analyzing (I) the

divergence of the weight average (M~) and the characteristic (M* ) molecular weights and of the radius of gyration (R~), and (iii the power law behavior of the time correlation function

and the divergence of the average relaxation time at the gelation threshold. In that way, the

scaling exponents were measured separately by different experiments for the same polymeric

system and can be compared with the theoretical predictions of their values and interrelations.

Experimental section.

The dicyanate used for the synthesis of the polycyanurate polymers, 2,2-bis(4- cyanatophenyl)propane I (dicyanate of Bisphenol A, DCBA) was synthesized with the help of the cyanbromide method [10]. The resins were prepared in bulk at 463 K in a stirred 100 ml flask which was heated in an oil bath. The reaction was stopped after reaching a highly viscous state by cooling the flask with liquid nitrogen. The same procedure was used for the synthesis

of the copolymers of DCBA with different admixtures of the monophenol 2-phenyl-2-(4- hydroxyphenyl)propane 2 (4-Cumylphenol, CP) (see Fig, I).

For measuring the time correlation function during gelation by online dynamic light scattering a DCBA-prepolymer (M~o

=

420 000 g/mol) at a concentration c

= 35 g/l was

placed into the measuring cell (2 ml) and 40 mg of the diamine 1,8-diamino-3,6-dioxaoctan

were added. The mixture gelled after ca, 12 h. A measuring time of 20 min for each run was taken.

The static and dynamic light scattering experiments were made simultaneously with

solutions of the polymers in tetrahydrofuran (THF) at 293 K with an automatic goniometer

and a structurator/correlator ALV-3000 by ALV-Langen. The red line (A

=

647 nm) of an

krypton ion laser Model 2000 by Spectra Physics was used as light source. The solutions were filtered before measuring through a millipore 5 ~Lm filter. The online measurements during gelation were done at a scattering angle of 90°.

(4)

~i"")

N/

OI

3n NEC-O-R-O-C+N -

~jS~

~ll~Q~/~~~~

,R~ 'R,

/~, ° ° /~@

l'~,,~ ~,J~

' n

~~~-CWN ~~~~~

i~~

NWC-°~j~~

~ ~

l 2

Fig, I. Scheme of the polycyclotrimerization of dicyanates and structures of the used monomers DCBA (I) and CP (2).

Gelation behavior static measurements.

Gelation of a trifunctional randomly branched polymeric system is known as a typical critical

phenomenon. In both the classical Flory/Stockmayer [5] and the three-dimensional perco- lation theory [I I] the gelpoint of a covalently crosslinked polymer is defined by the divergence of the second moment of the molar mass distribution, but, on the other side, a different

critical behavior is predicted [I Ii.

Pure DCBA-based polycyanurates were shown to have a critical conversion of the cyanate group at the gelpoint of 50 fb for the reaction in bulk [3], which exactly coincides with the

prediction of the classical theory [12, 13]. Unfortunately, the required accuracy, which is needed to calculate critical exponents, can not be achieved by spectroscopy as pointed out in

our previous paper [Ii. To overcome these difficulties the polymeric system was slightly

modified by admixing the monofunctional phenolic analogon CP to the difunctional cyanate DCBA. In this way, the critical conversion of the OCN-group can be replaced by the critical initial fraction of hydroxyl groups, since the mixture forms a gel at full conversion below this critical value only. On the other side, completely soluble polymers with high mean molecular

weights can be synthesized, if one uses mixtures containing more monofunctional units than the critical fraction. The determination of the distance to the critical value is now a weighing

process, which can be done with a higher accuracy than the conversion measurement by

infrared spectroscopy.

The mean molecular dimensions of the polymers are only negligibly influenced by the

admixtures of the monofunctional monomers. This is demonstrated in figure 2 where the radii of gyration of pure polycyanurates and of the copolymers are plotted versus the weight

average molecular weight. All polymers follow the same scaling relation. Furthermore, from measurements with the help of size exclusion chromatography coupled with light scattering

(5)

iO~

E

~ iO~

io~

iO~ iO~ iO~ iO~ iO~

h4w

Fig. 2. Radius of gyration R~ versus molecular weight M~ for pure DCBA- (m) and DCBA/CP- (D) polycyanurates. The regression line follows relation (1).

jQ8

bflmw iQ~

jQ6

~/W

jQ5

lo~~ lo~~

c

Fig. 3. Increase with decreasing relative distance e to the gelation threshold of the mean molecular

weight M~ (m) and the molecular weight M~~ (D) corresponding to the maximum of the light scattering

curve of the GPC/LALLS chromatogram. The regression lines follow the relations (3) and (4).

the same power law for the molecular weight distribution is found as shown by figure 3 in our

previous paper [I ].

The divergence of the weight average molecular weight at the gelation threshold is shown in

figure 3. Here, s

= (X-X~)/X~ is the relative distance of the normalized initial molar fraction of hydroxyl groups X

= [OH ]o/([OCN Jo + [OH Jo) to the critical value X~. The

optimum values of X~ and y, which are highly correlated, are calculated by an iteration procedure minimizing the residual mean deviation of the experimental points to the

regression line as illustrated in figure 4. In that way, a scaling law

M~

~ s~ ? with y

= 1.7 ± 0.04 (3)

was obtained, where the exponent y is in very good agreement with the three-dimensional

percolation prediction but remarkably differs from the Flory/Stockmayer value of 1.

(6)

-1 o-lo

I O.QB

i ',,,, ~

,, j

"", ,,'~ O.OfS )

~ _~ ',,, ,,,

",,, ,/" O.04 c

E ',,,~ _,,' [

)

O.02

~

-3 O.OO

24.O 24.1 24.2 24.3

Xc

Fig. 4.- Dependence on the choice of the critical initial molar fraction of hydroxyl groups X~ of the slope y defined by relation (3) (-) and of the mean deviation between experimental and

calculated points (----). The range of the optimum value of y corresponds to the range of the minimum of the deviation between the dotted lines.

Furthermore, the divergence of the molecular weight M~~ corresponding to the maximum of the light scattering curve of the GPC/LALLS chromatogram is shown in figure 3. This

value was shown by Schosseler et al. [14] to be proportional to the cutoff molecular weight M*, the divergence of which defines an exponent «. The experimentally found relation :

M~~~

~ s

~'" with «

=

0.52 ± 0,12 (4)

lies near the predictions of the classical

=

0.5 ) as well as the three-dimensional percolation theory

= 0.46) but its error level is higher than the difference between both theoretical values so that a clear distinction is not possible with the help of « alone.

On the other side, the exponent T, which describes the width of the molecular weight

distribution (2), should be directly calculated from the M~~~ dependence of M~ as pointed out

by Schosseler et al. [14]. The least squares fit through the data shown in figure 5 gives the relation :

M~

~

M(£~ with T

=

2,18 ± 0.05 (5)

This value agrees very well with that calculated directly from the slope of the molecular

weight distribution.

The experimental data which show the divergence of the mean cluster size R~ at the gelation

threshold are plotted in figure 6 together with the regression line

R~ s~ ~ with v =

0.9 ± 0.08. (6)

The exponent v also coincides with the three-dimensional percolation prediction (v

= 0.88) [1Ii and lies significantly higher than the classical value (v = 0.5 ). In that way, the basic relations between the different exponents [I Ii are fulfilled using our experimental

values :

p = (T 1)/3

=

0.41 p~~~ =

0.41 ± 0.02 (7)

p = Ma = 0.38 (8)

« =

(T 1)/3 v

=

0.45 «~~~ =

0.52 ± 0.12 (9)

« = (3 T)/y

=

0.46 (10)

(7)

iQ~

~~6

~~5

~~4

iO~ iO~ iO~ iO~

~Amax

Fig. 5. Mean molecular weight M~ versus molecular weight M~~ corresponding to the maximum of the light scattering curve of the GPC/LALLS chromatograrn. The regression line follows relation (5).

iQ~

~ ~~2

~~~

io~~ io~~

Fig. 6. Increase with decreasing relative distance e to the gelation threshold of the radius of gyration R~. The regression line follows relation (6).

Gelation behavior dynamic light scattering.

It has been shown previously that for covalently as well as for physically crosslinked networks both the power law behavior of the time autocorrelation function (TCF) of the scattered electric field and the divergence of the characteristic relaxation time are criteria for the

gelpoint [15-19]. The TCF is obtained from dynamic light scattering [20] and may

approximately be regarded as a Laplace transform of the relaxation spectrum in the observed system. For solutions of polydisperse macromolecules having a broad distribution of

relaxation times the corresponding TCF can be written as a superposition of exponential decays

gi (t)

= j~ h( Tj exp (- t/Tj dT (I I)

o

(8)

where h(T ) is a distribution function. Such a superposition can often be well approximated by

a so-called stretched exponential function [21, 22]

hi (ti

= a exp i- (t/bi~i + d (121

Here, a is the correlation strength, b is related to the mean relaxation time (T) of the spectrum through

jTj

= (b/cj r(i/cj (13j

where r is the common gamma function, c contains inforrnations on both the mass

distribution of the polymer and the hydrodynamic interaction, and d describes a possible non- zero baseline.

In a previous paper [19] we introduced a hybrid of a stretched exponential and a power law

gi(t)

= hi(t) h~(t) h~(t)

= (tie)~f~~~~ C (t) = I exp(- t/g) (14)

to describe the TCF in the whole range from the pregel through the gelpoint into the gel state with just one single type of function. Here, hj(t) is a stretched exponential defined by equation (12) which approximates on the one side the whole TCF far away from the gelation

threshold and on the other side the initial fast decay, if the power law works in the vicinity of the gelpoint. The function h~(t) is a power law with a simple cutoff C (t) of the exponent that guarantees a continuous and smooth transition from the power law behavior to the stretched

exponential decay which is valid at short times. In the same way, the parameter

e rescales the time so that both parts hi (t) and h~(t) of the hybrid form a uniform function.

A selection of time correlation functions measured during the chemical gelation of the

polycyanurate in THF is shown in figure 7. From this double logarithmic plot it is easy to see

qualitatively that the TCF goes from a completely decaying behavior during the pregel state

through a power law over a wide range of times at the gelpoint into a non decaying form with

a non zero plateau in the gel state. It should be noted that the same kind of curves was obtained by Lang et al. [18, 19] during the thermoreversible gelation of a polysaccharide.

iO°

gel

a

increasing j

~ ~o-i reaction time *~

~ °

°a I

~ %

o m

pregel o jo

&

$(

~~-2 °

iO~~iO~~iQ~~iO~~iO~~iO~~lO~~ iO° lO~ iO~

t / s

Fig. 7. Dependence on the reaction time of time correlation functions gj(t) of a polycyclotrimerizing polycyanurate in THF. The measured curves correspond to reaction times of 14 and 627 mn (D), 673 mn (m), 721 and 821 mn (+). The function gj (t) shows a power law behavior over four decades in time with

a slope of f = 0.34 in the vicinity of the gelpoint (m).

(9)

With the help of the model function (14) all measured curves could be well fitted and the

development with increasing reaction time of the optimum parameters is shown in the

figures 8a-c. As one can see, all parameters show a characteristic course, what again

demonstrates that they are not arbitrary fitting parameters but have a physical or

mathematical significance connected with the kind of curves under consideration.

a, d b, e, Q

i-o io-~

iQ-2 Q ,l'~

',, ---.---'~----.-~-fi~W(.~~W~.

°.~ ~"., iQ~3

Q ~

a "',,

b ' "

',.

"'~'~"~,~ lO~~ ')

~'~

~ i

lo' h

Q4 Q--a"~"~' lO~~ '~

,°"

d ,,'°'

lo-?

,Q" /

02

lO~~ e

a

OO i

O 300 SOO 900 O 300 SOO 900

reaction time / min reaction time / min

a) b)

C >~

i o

off

OA

~

/,,(

, ',

o ~ / a

~ ,"~ "~~

"D o-o

o 300 aoo goo

reaction time / min

C)

Fig. 8. a-c) Dependence on reaction time of the optimum fit parameters according to equations (12) and (14).

The correlation strength a decreases slowly during the initial part of the reaction and reaches a significantly lower value in the vicinity of the gelpoint. The opposite behavior is, of

course, obtained for d, which represents the contribution of the power law and, after the

gelpoint, the gel fraction respectively. The non,zero value of d at the beginning of the reaction is due to the experimental noise of the baseline.

The characteristic time b, from which via equation (13) the mean relaxation time of the fast motions can be calculated, as well as the appropriate exponent c remain nearly constant

during the whole range of reaction times. This behavior indicates that the translational

diffusion with strong hydrodynamic interaction (ZIMM modes with an expected exponent

(10)

c = 2/3) [23] of the small particles inside the solution is not significantly influenced by the

formation of the macroscopic gel during the crossover from the semidilute regime and

M~

=

420 000 into the gel state including a large fraction of sol.

The increase of the mean relaxation time of the whole system during the reaction progress up to the gel point is reflected by the increasing effect of the power law term

h~(t) what is manifested in an increase of d, e and f. Especially, the sharp increase of the exponent f and the rescaling time e in the vicinity of the gelpoint is a quantitative measure for the power law behavior over some decades in relaxation time being a characteristic of the gel point. This fact is substantiated by the behavior in the gel region, where the power law

fraction now decreases with advancing reaction time but the gel fraction increases,

It must be noted that the slope of the linear part in the double logarithmic plot of the TCF at the gel point f= 0.34± 0.02 is the same as found for a thermoreversibly gelling polysaccharide [19] and lies close to the value reported on semidilute solutions of

polyurethane clusters [16] (f= 0.23-0.38). However, it is two times higher than the

experimental value for covalent silicon gels [17] ~f= 0.27 for the homodyne TCF g(t)~ [gj(t)]~). Unfortunately, the long time tail of the correlation function was not

sufficiently accessible on the experimental time scale. Therefore, an analysis of the divergence

of the slowest characteristic time was not possible with acceptable accuracy.

The scaling behavior of the mean relaxation time (T) of the spectra at the gelation thresh- old is shown in figure 9. Here, (T) was calculated by an integration of the TCF over the

whole experimental time window

j~j

=

~~ gj(t) dt (15)

and s

= (t~ t)/t~ represents the relative distance to the critical reaction time t~. From a

regression analysis done in the same manner as described for the calculation of X~ and y (see Fig. 4) the relation

(T) s~~ with #i = 1.9 ± 0.2 (16)

was obtained. The same exponent was found by Martin et al. [17] for covalent silicon gels but through an integration of the homodyne TCF.

io'

m iO°

) iO~'

v <~>

lO~~

<~>f

u

~Q-3 m

io~~ io~' io°

c

Fig. 9. Mean relaxation time (T ) of the whole function (m) and (T)~ of the fast motions (c) versus relative distance e to the critical reaction time. The regression line for the divergence of

(T) follows relation (16).

Références

Documents relatifs

L’archive ouverte pluridisciplinaire HAL, est destinée au dépôt et à la diffusion de documents scientifiques de niveau recherche, publiés ou non, émanant des

La méthodologie dans ce projet repose sur le principe de la modélisation markovienne (modèle de la progression du CHC sur VHC) pour évaluer l'impact du dépistage du CHC en termes

One-sided exact boundary null controllability of entropy solutions to a class of hyperbolic systems of conservation laws. Weak and measure-valued solutions to evolutionary PDEs,

Ces données sont en accord avec celles obtenues in vivo, puisque la micro- injection de morphine dans la VTA n’induit plus d’augmentation de l’activité électrique des

L’archive ouverte pluridisciplinaire HAL, est destinée au dépôt et à la diffusion de documents scientifiques de niveau recherche, publiés ou non, émanant des

Maekawa, On asymptotic stability of global solutions in the weak L 2 space for the two-dimensional Navier-Stokes equations, Analysis (Berlin) 35 (2015), no. Pileckas, and R

Le ministre chargé de la santé ainsi que l'Agence nationale de santé publique, un organisme d'assurance maladie et les agences régionales de santé peuvent en outre, aux mêmes fins

A systematic study of the physical and chemical behavior of mineral matter during rapid combustion of pulverized coal particles was performed by first