Light Scattering of PMMA Latex Particles in Benzene: Structural Effects
Transcription
Light Scattering of PMMA Latex Particles in Benzene: Structural Effects
Light Scattering of PMMA Latex Particles in Benzene: Structural Effects E. A. NIEUWENHUIS AND A. VRIJ Van't H o f f Laboratory for Physical and Colloid Chemistry, Transitorium 3, Padualaan 8, 3584 CH Utrecht, The Netherlands Received February 19, 1979; accepted February 19, 1979 Intra- and interparticle structural effects were studied in polymethylmethacrylate (PMMA) latex dispersions in a nonpolar solvent with the technique of light scattering. The required transparency of the dispersions was attained by a close matching of the refractive index of PMMA and solvent, for which benzene was chosen. Two latices were studied with divinylbenzene (DVB) and ethylene glycol dimethacrylate (EGDM) as crosslinking agents. The latex particles were characterized by electron microscopy in the dry state and by viscosity, sedimentation, light scattering, and laser light scattering spectroscopy in dilute solutions. The molar mass of the particles is in the range of 5 × 109 g mole-L In benzene they are swollen and a hydrodynamic radius in the range of 200 nm was measured. The light scattering intensity and its angular variation were found to be strongly temperature dependent, This could be explained by the difference in temperature coefficients of the refractive indexes of PMMA and benzene which is sufficient to change their matching appreciably. For the latex containing DVB this phenomenon could be used to characterize the spatial distribution of the DVB inside the particle. At higher concentrations (20-80 g dm 3) light scattering as a function of scattering angle shows the emergence of a diffuse diffraction peak which points to interparticle structural effects. In one case the radial distribution function of particle centers was determined by a Fourier transformation of the structure factor and is discussed in terms of some theories of atomic liquids. The position of the diffraction peaks as a function of concentration is also discussed in terms of an "expanded-lattice" structure, Results suggested that some structure persisted down to concentrations of 2% (w/v) of latex. However, the calculated mean interpa~icle distances seem much too large to be explained on a basis of interpenetration of peripheral chains of the particles. Attempts were made to interpret the long range interaction by electrostatic repulsions. However, measurements of the electric conductivity of the dispersions do not corroborate this view. Similar structural effects showed up with laser light scattering spectroscopy, where it was found that the diffusion coefficient depended on the scattering angle. On prolonged standing some of the systems showed sharp, bright diffraction peaks in visible light due to the formation of "supramolecular crystals" as reported earlier in the literature. few fundamental investigations have been reported, the more recent ones being those of Krieger et al. (2) and Hachisu et al. (3). At higher concentrations these dispersions show the typical iridescence of visible light caused by interference of light waves scattered by the individual particles which are packed in a regular fashion. A serious drawback in studying this iridescence is the very large optical density of these dispersions. This restricts the scope of these studies to thin surface layers, the properties of which 1. INTRODUCTION In 1958 Shashoua and Beaman (1) published the first systematic work on dispersions of polymer latex particles in organic solvents. Already at that time they stressed the importance of these systems for studying fundamental aspects of polymer behavior. The individual particles were called microgets. The dispersions were characterized by ultracentrifugation, viscosity, and electron microscopy. Since then relatively 321 Journal of Colloidand InterfaceScience, Vol. 72, No. 2, November1979 0021-9797/79/140321-21 $02.00/0 Copyright © 1979by AcademicPress, Inc. All rightsof reproductionin a~lyform reserved, 322 NIEUWENHUIS AND VRIJ TABLE I PolymerizationData of Latices Latex Concn of MMA (mole dm -3) Concn of KPS (mole dm -3) Ratio crosslinking monomer (vol%) Reaction time (hr) Reaction temperature (°C) P10 P12 1.0 1.0 9.25 × 10-4 3.33 × 10-4 EGDM (5%) DVB (5%) 7 8 70 70 may be influenced by the presence of the wall of the vessel, it would therefore be of interest if this large optical density could be alleviated for in that case one could also study the scattering properties of the bulk of the dispersion, just like X-ray scattering in the case of molecular liquids, and obtain information about the internal structure of the particle system. In order to investigate such a possibility we have chosen to study crosslinked PMMA latex in benzene. In these dispersions the refractive index difference between particles and solvent is very small and the systems are transparent. In this paper we will present results on the average light scattering intensity as a function of scattering angle and dispersion concentration as well as some results on the time correlations in the scattering intensity (quasi-elastic light scattering). The first method gives information on the time average structure in space and the second method on spatial structure changes with time. Both kinds of information are essential ingredients in the formulation of better theories on concentrated colloidal dispersions in general. In the near future one could hope to build such theories on the foundations laid by the theories of simple, molecular liquids, which have improved so much in the last 10 to 15 years (4). The dispersed particles, acting as "supramolecules," embedded in a "structureless" low-molecular-weight solvent, would play the role of the molecules in a simple, monoatomic liquid (5, 6). Journal of Colloid and Interface Science, Vol. 72, No. 2, November 1979 2. LATEXPREI:;ARATION Following Ono et al. (7) the latex dispersions were prepared by heterogeneous polymerization in water without emulsifier using different amounts of potassium persulfate (KPS) as initiator. This method, yielding so-called "soap-free" polymer latex dispersions, was adapted by adding the crosslinking agents divinylbenzene (DVB) (3) or ethylene glycol dimethacrylate (EGDM) to the reaction mixture. The latter (EGDM) has a refractive index nearly equal to that of MMA, thus yielding an optically more homogeneous particle than DVB which has a higher refractive index. The polymerization conditions of the two latices are listed in Table I. It turned out that without further purification the vacuum-dried latex did not disperse in benzene. Therefore the latex was purified by dialysis (Union Carbide tube) against deionized water for a few days to remove residual MMA. During this period the dialysate was changed continuously. Excess initiator and other ionic impurities were removed by ion-exchange (7-9), using equal weights of Dowex 50W-X4 and Dowex l-X4 resin. The purified latex was vacuum dried at room temperature and was readily dispersable in benzene under shaking in a few hours. The solutions were almost transparent when viewed under diffuse light. Finally, excess free polymer was removed by sedimentation of the latex particles and redispersion in fresh solvent. Repeating this procedure three times proved to be sufficient. 323 STRUCTURE IN LATEX DISPERSIONS 3. INSTRUMENTAL The light scattering (LS) measurements were performed with a F I C A 50 light scattering photometer. As a scattering standard pure benzene was used (R90 = 15.8 × 10-6 cm -1 at )to = 546 nm andRa0 = 45.6 x 10-6 cm -1 at )to = 436 nm). The angular range studied was 15° ~< 0 ~< 150 °. The incident light beam was unpolarized. The solutions were filtered through Millipore filters (1.2 and 3/xm) directly into the measuring cells. Laser light scattering spectroscopy (LLSS) was performed with a laboratory-built instrument containing a double-walled thermostated sample holder. The 514.5-nm line of an argon ion laser (Spectra Physics Model 165) was used and the light was detected with an E M T 9558 photomultiplier. The correlation functions of the fluctuations in the scattered intensity were determined with a S a i c o r - H o n e y w e l l 42 A photon correlator with p h o t o n counting option. The correlation functions were numerically analyzed. For measurement of the specific refractive index increments a B r i c e - P h o e n i x differential r e f r a c t o m e t e r was used. Transmission electron microscopy measurements were performed with a Philips EM 301 apparatus. Carrier grids c o v e r e d with carboncoated Parlodion films were dipped in a dilute dispersion and micrographs were taken of the particles retained on the film. Viscosity measurements were carried out with Ubbelohde internal dilution viscosimeters at t = 25.0 _ 0.01°C. Sedimentation velocity experiments were carried out in a Beckman Spinco (Model E) analytical ultra- FIG. 1. Electron micrograph of crosslinked PMMA latex particles (sample P12) as taken from the aqueous suspension after the preparation. centrifuge using a 3-cm centerpiece and an AN-J rotor at 25°C. Rotor speeds were used between 2000 and 3000 rpm. 4. PHYSICAL CHARACTERIZATION OF LATEX a. Electron Microscopy Figure 1 shows latex particles as taken from the aqueous phase. The particles are almost perfect spheres with a standard deviation in the radius, aEM, of about 10%. The found absolute values of the average radii, as given in Table II, should be considered with caution, however, because it was observed that shrinkage and fusion of the PMMA particles occurred at higher intensities of the electron beam. Similar p h e n o m e n a were reported by Fitch (9) for TABLE II Physical Characterization of Latex Particles Latex aEM (H20) (nm) OD (H20) (nm) aG (H20) (nm) Do (benzene) (10-Scm2sec i) aB (benzene) (nm) S0 (benzene) (10 ~sec) k~ (cmZg-I) M (10~g mole -1) [7] (g l c m z) a~ (nm) P10 PI2 103 115 130 161 129 158 1.88 1.38 177 240 1.03 1.18 54 33 4.6 7.1 8.8 9.4 210 265 Journal ofColloM and ~terface Science, V01.72, No. 2, November1979 324 NIEUWENHUIS AND VRIJ PMMA, and by McDonald (!0) for PVC, particles. b. Laser Light-Scattering Spectroscopy The correlation function of the fluctuations of the scattered light was measured from dilute latex dipersions in water as well as in benzene. In both solvents the experimental data showed a single exponential decay. According to the theory, as described in detail by Berne and Pecora (11), the autocorrelation function, F(K,r), for the fluctuations in the scattered intensity of noninteracting particles is given by F(K,.r) ~ exp(-2DoK2z). [1] Here Do is the diffusion coefficient, K = (47rn/ho) sin (0/2) is the wave vector of the fluctuation, with n the refractive index of the medium and ~o the wavelength of the light in vacuo, and z is the correlation time. By plotting log F(K,.r) against z for a certain K a straight line was obtained. The slopes were plotted against K 2, giving a straight line the slope of which gives Do according to [1]. For dilute dispersions the hydrodynamic radius of the particles, aD, is related to Do by the Stokes-Einstein law, Do = kT/(67r*/oaD), [2] is observed during the experiment. The sedimentation coefficient s was plotted according to the relationship s -a = s~-l(1 + k~c). Here So is the sedimentation coefficient at infinite dilution and k~ is a proportionality constant (12). From So, Do, and 1 - v~0, the molar mass M of the particles was calculated (see Table II). Here ~ is the partial specific volume of PMMA (13): ~ = 0.807 cm ~ g-l, and ~0= 0.874g cm -z is the density of the solvent benzene at 25°C. d. Viscosity The relative viscosity */r was measured and the specific viscosity ~/sp = (*/r - 1)/c was then calculated for P10 and P12 in benzene at 25,0°C. Most of the measurements were performed on samples that still contained some free polymer. The measured data were corrected afterward, simply by considering the supernatant after centrifugation as the continuous phase instead of pure benzene. Some checks on pure samples gave identical results. Intrinsic viscosities [*/] = lim (c = 0)*/sp were obtained by extrapolation of */sp to zero concentration according to the equation (3), */sp = (5/2)q + Eq2c. [3] where k is Boltzmann's constant, T is the absolute temperature, and */0 is the viscosity of the medium. The so obtained diffusion coefficient Do appears to be only slightly dependent on the concentration of the latex in dilute systems. Final results are sum. rnarized in Table II. Here q is a scaling coefficient to fit the term linear in c to the Einstein equation: q = (2/5)[,/]. It can be expressed as q = o~b, where c~ is a volume expansion factor. A hydrodynamic radius an can be calculated from a n = a~aD(H20). Both [*/] and an are given in Table II. The results are comparable to those obtained earlier (3). c. Sedimentation e. Light Scattering Because of the small difference in refractive index between particle and solvent Schlieren optical measurements in benzene could be done only above concentrations of ! g dm-L The Schlieren picture was a single sharp peak; only a small peak broadening We measured the angular dependence of the light scattered from very dilute aqueous dispersions. The results were analysed in a so.called Guinier plot, in which the logarithm of the scattered intensity (I) is plotted against the wave vector Journal of Colloid and Interface Science, Vol, 72, No. 2~ November t979 325 S T R U C T U R E IN L A T E X D I S P E R S I O N S (K) squared. According to Guinier (14) the scattered intensity at low angles (in the R a y l e i g h - G a n s - D e b y e approximation) falls off with the wave vector as I ( K ) ~ exp x (-K2a~/5) for small K, where aG is the radius of the latex sphere (in water). The radius of the latex particles was obtained from the slope of the straight part of the Guinier plot at small K and is given in Table II. The agreement with radii obtained from the diffusion experiments is v e r y good. The p r o c e d u r e is based on the R a y l e i g h G a n s - D e b y e approximation which is valid when m, the ratio of the refractive indices of particles and solvent, is near unity. For other values o f m the theory of Mie must be used. H o w e v e r , according to K e r k e r (15) the difference between both theories should be less than 10% for our case where m = 1.12 and the sizes are still relatively small. So we have used the much simpler RGD approximation. [ R~(K)/c]/cm 2 g-1 io I t- I0 o 10 10 10-Jl 0 , , 4 - - m ~ - - - - ~ 8 K 12 cm FIG. 2. Variation of light scattering intensity with scattering angle (0) at different temperatures for latex sample PI2 in benzene. K = (47m/h0)sin (0/2). ~0=546 nm; c ~ lg dm-3. Data are given in a Guinier plot for 21°C (O), 17.5°C ([]), I0.5°C (©), and 6.5°C (11).R*(K) = R(K)/(1 + cos2 0). 5. LIGHT S C A T T E R I N G OF D I L U T E D I S P E R S I O N S IN B E N Z E N E ; EFFECTS OF PARTICLE STRUCTURE a. Light-Scattering Equations Light scattering is one of the available techniques to obtain structural and thermodynamic information on dispersions. F o r a monodisperse system of spherically symmetric particles one may write for unpolarized light in the R a y l e i g h - G a n s D e b y e approximation (16), R ( K ) = (1 + cos 20)Y{cMP(K), [4] with K = (47m/X0) sin (0/2). Here R ( K ) is the reduced scattering intensity (Rayleigh ratio) of the dispersion o v e r that of the solvent, 0 is the scattering angle, M the particle molar mass, P ( K ) = Po is the particle scattering factor, and c = o M / N A is the weight concentration with p the particle n u m b e r density. Further, if{ = 2 7 r 2 n 2 ( d n / d c ) 2 ( X ~ N h ) -1. [5] Here n is the refractive index of the dispersion, X0 the wavelength of the light in vacuo, and NA Avogadro's number. b. Experiments Because the particle mass is v e r y large, a solvent (benzene) was chosen in which the refractive index increment dn/dc of the particles is very small. At 20.0°C we found for P10: d n / d c - ~ 0.0075 (X0 = 546 rim); 0.0009 (X0 -- 436 nm) and for P 1 2 : 0 . 0 1 1 0 (X0 = 546 nm); 0.0048 (X0 = 436 rim), all positive, and in units cm~/g. The accuracy is low because of the limited sensitivity of the refractometer. This implies that it is not feasible to obtain a good absolute value of M from light scattering intensity. Note that dn/dc of P12, containing DVB as a Journal of Colloid and Interface Science, Vol. 72, No. 2, November 1979 326 NIEUWENHUIS AND VRIJ tRIKI/cl/J g-1 I01F 100 I(]I 16 2 1(]~IO i ~ t j 8 i i 12 I I I 16 2 r K cm, FIG. 3. As Fig. 2 but for k0 = 436 nm. cross-linker, is larger than that of P10, which contains E G D M . It turned out that the light scattering was rather sensitive to variations in temperature. Results for X0 = 546 and 436 nm for P12 are shown in Figs. 2 and 3. One observes that R ( K ) at small angles varies with a factor of about 3 for h0 = 546 nm and with a factor of about 50 for h0 = 436 nm. Also the slopes are temperature dependent, in particular for )~0 = 436 nm. It will be clear that the main cause of this abnormal behavior must be searched in (small) changes of dn/dc with T. In most cases these small changes are not observable. In our systems, however, the absolute values of dn/dc are so small that these changes may b e c o m e of comparable magnitude. This would still have no consequence on the shape of the light scattering curves if the particle compositions were completely homogeneous. It will be apparent, h o w e v e r , that any deviation of homogeneity, e.g., because of the Journal of Colloid and Interface Science, Vol. 72, No. 2, November 1979 presence of a small n u m b e r of crosslinks with a rather different dn/dc, might also change the shape of the curves. In turn, these changes in shape will then contain information about the inhomogeneity. This will now be investigated in more detail. c. Scattering o f Particles with Several Scattering C o m p o n e n t s We assume that the swollen P M M A latex particles contain two scattering components, i.e., P M M A chains with a refractive index increment dn/dc = vB and a relatively small n u m b e r of crosslinks, e.g., DVB, with a dn/dc = va. It will be apparent that such a particle is in fact a degenerate case of a c o p o l y m e r molecule with two types of chains, the scattering behavior of which is known (17). The swollen latex particles, however, have some features often not encountered in the more familiar copolymers. The particles are practically spherically STRUCTURE -Y2 s y m m e t r i c and h o m o g e n e o u s in size ( s o m e d i s p e r s i o n s o f these particles f o r m supram o l e c u l a r crystals; see S e c t i o n 7). T h e centers o f m a s s o f b o t h units A and B will n e a r l y coincide. Light scattering near 0 = 0. F o r 0 --~ 0, P ( K ) in [4] b e c o m e s u n i t y and for particles p o l y d i s p e r s e in size and c o m p o s i t i o n , vZM m u s t be r e p l a c e d b y (17) v2M ~ ~ TiMiu ~ . 327 IN LATEX DISPERSIONS 3 [6] i H e r e v = the m e a s u r e d dn/dc o f the particles; y~ = q~ ~ c , M~ is the m o l a r m a s s o f species i and 5 10 15 20 t:/% l) i -~- W i l ) A "~- ( 1 -- Wi)IIB, [7] where w~ is the weight fraction of the crosslinker (A) in species i. F u r t h e r [6] m a y be written as F r o m Eq. [7] one has with wi = w v2M = v~Mw + 2PV(VA + VB) + Q(vA FIG. 4. The square root of the scattering at zero angle divided by the concentration as a function of the temperature for sample PI2 in benzene. (3, X0 = 546 nm; ~, X0 = 436 nm. R*(0) = R(0)/2. - v~) ~, [8] d v M T = wdvAMT + (1 - w)dvBMT. [10] F o r the s e p a r a t e c o m p o n e n t s we take (17) where P = ~ yiMi(w~ - (v), re= be(he-no), i Q = ~ yiMi(w~ - vb)z, [9] ( ( = A , B), [11] w h e r e be is the partial specific v o l u m e and i and M is the a p p a r e n t and Mw the weight a v e r a g e m o l a r m a s s o f the particles. B e c a u s e it c a n be e x p e c t e d that v is a linear f u n c t i o n o f t e m p e r a t u r e w e h a v e p l o t t e d [VzR(O)/c] 1/z as a f u n c t i o n o f T as s h o w n in Fig. 4. T h e plots are linear, so a p p a r e n t l y the terms c o n t a i n i n g P and Q do n o t c o n t r i b u t e p e r c e p t i b l y . F r o m Fig. 4, w e calculated v as a f u n c t i o n o f T as s h o w n in Fig. 5. We u s e d M = 6.0 x 10~g m o l e -1 to obtain a g o o d fit with the t w o values o f v m e a s u r e d with the r e f r a c t o m e t e r . T h e M obtained from sedimentation-diffusion m e a s u r e m e n t s is 7.1 x 109g m o l e -~ (see Table II) thus in r e a s o n a b l e a g r e e m e n t . L e t us n o w investigate h o w far the t r e n d with t e m p e r a t u r e c a n be explained on the basis o f separate variations in vA and vB. 9/10 -3 c m 3 g-I 12 / J J y J J 0 5 1~0 1~5 210 ~/°c Fio. 5. Specific refractive index increment, v, as a function of the temperature, t, as calculated from the light scattering results assuming M = 6.0 × 109g mole -~ for sample P12 in benzene. (3, h0 = 546 rim; A, h0 = 436 rim. Also indicated are two measured values of v (11). Journal of Colloid and Interface Science, Vol. 72, No. 2, November 1979 328 NIEUWENHUIS AND VRIJ [ rg 2//3] /I011 cm 2 6 2 0 t 0 5 I 10 15 , 20 ~ 25 -1 ( t - t ' ] -1/10 -2 K FIG. 6. Dependence of the apparent radius of gyration, rg, on temperature for sample P12 in benezene at X0 = 546 nm (O) and )to = 436 nm ([~). Note that T* (T* = T at which dv/dt = 0) is different for both wavelengths (see text). ne is the refractive index of c o m p o n e n t and no the refractive index of the solvent. This equation reflects the influence of the solvent on re. Using this equation for a single p o l y m e r in different solvents it was found (18) that the variation of v~ with no is indeed linear and that ne is v e r y near to the refractive index of the bulk p o l y m e r . Then one m a y write dvjdT P(K) = exp[-K%~/3]. [12] F o r ve and d ( n , - no)/dT we will take values for bulk p o l y m e r and solvent (17, 19). T h e n one finds for A (crosslinker): 104 dvA/ d T = +4.8 c m 3 g-1 K-1 and for B (PMMA): 104 dvBMT = +2.8 c m 3 g-1 K -1, independent of wavelength. Thus 104 dv/dT = 4.8w + 2.8(1 - w) = 2.8 + 2.0w. This is rather independent of the (small) value of w. The values found f r o m the light scattering exp e r i m e n t s in Fig. 5 are 104 d v M T = 3.3 cm 3 g-1 K -~ for h, = 546 nm and 3.1 cm 3 g-~ K -~ for X0 = 436 nm. The wavelength i n d e p e n d e n c y is in a c c o r d a n c e with theory. The calculated values for dv/dT h o w e v e r are s o m e w h a t small. Journal of Colloid and Interface Science, Vol. 72, N o . 2, N o v e m b e r 1979 [13] F o r a c o p o l y m e r the radius of gyration, rg, takes the f o r m (17) r2g = r2gB + y(r~A-- r2gB) + y ( I -- y ) F , = f~d(n e - no)/dT + v~ldb~/dT. Radius o f gyration. L e t us now investigate the light scattering b e h a v i o r at finite scattering angles as o b s e r v e d in Figs. 2 and 3. F o r the linear parts of the plots one has the Guinier a p p r o x i m a t i o n (14) for P ( K ) : [14] with y = WVA/V. H e r e rg2A and r~B are (average) square radii of gyration of the c o m p o n e n t s A and B and F is the (average) square distance bet w e e n the centers of A and B. N o t e that according to Eq. [14] rg depends on v and is often called a p p a r e n t radius of gyration. In our case y takes the form, y(T) = WVA[(dv/dT)(T - T*)] -1, [15] where T* is the (extrapolated) t e m p e r a t u r e where v ~ 0. F r o m Fig. 5 one obtains T'a6 = 259.3°K and T*36 = 275.3°K. Equation [15] suggests to plot the slope of the straight part of the natural logarithm of P ( K ) versus K 2 as a function of ( T - T*) -1. The result for 329 S T R U C T U R E IN L A T E X D I S P E R S I O N S R~K) ~10-2 cm -1 6.4 4.8 3.2 1.6 I ~ 2 i~ 4 6 8 ,,, ,± 10 K2 / I 12 1010 cm-2 FIG. 7. R e d u c e d light scattering intensity, R * ( K ) , v e r s u s K z for latex sample P12 in b e n z e n e at h0 = 546 n m , as m e a s u r e d at a t e m p e r a t u r e of 6.5°C. T h e concentration (c/10 -2 g cm 3) data are as follows: (11), 1.36; (D), 2.26; (A), 3.77; (O), 5.39. R * ( K ) = R ( K ) / ( 1 + cos 2 0). P12 is shown in Fig. 6. The values for X0 = 436 and 546 nm both fall on the same straight line. Apparently the term in Eq. [14] causing nonlinearity is negligible in our case. The extrapolated value at (T - T*) -~ --> 0 gives rgB = 145 nm. Because vA and d v / d T are positive the negative slope implies that rg2A< r~B. Indeed, one would e x p e c t the spatial distribution of the crosslinker (DVB) to be less extended than that of the PMMA. Unfortunately the values of w and VA are not well known. Because the initially used w in the reaction was - 5 % , its actual value in the particle cannot be larger. The estimated value of uA was ~0.11 cm3/g. The product w vA can be estimated from the difference v(P12) - v(P10) ~ 0.0035 for X0 = 546 nm and 0.0039 for X0 = 436 nm (see [7]), Taking w u A = 0.0037 as a c o m p r o m i s e one finds from the intercept on the abscissa of Fig. 6: rgA/rg B --~ 0.8 or rgA ~ 115 nm. Further one finds w -~ 0.034 which seems reasonable, 6. L I G H T S C A T T E R I N G A T H I G H E R CONCENTRATION; 1NTERPARTICLE STRUCTURAL EFFECTS The relatively weak scattering power of the particles makes it possible to investigate these systems also at higher particle concentrations, Experiments were performed both with P10 and P12 at several temperatures. As an example, scattering curves o f the sample P12 at k0 = 546 nm are shown in Fig. 7 for T = 6.5°C where the overall scattering power is lowest (see Fig. 2). One observes that with increasing concentration well-defined peaks develop superimposed on the scattering o f the individual particles indicating the emergence of some order in the system. Similar curves were found for the sample P10. This behavior which is well known from X-ray scattering of simple liquids was observed and discussed earlier for small-angle X-ray scattering o f biological macromolecules by Riley and Oster (16, 20). It was also found recently in light scattering o f very dilute dispersions of latex particles Journal of ColloM and lnterfoee Science, Vol. 72, No. 2, November 1979 330 NIEUWENHUIS AND VRIJ in deionized water (21, 22) where it is caused smaller magnitude and a smaller angular deby extraordinary long-range interactions of pendence in the lower K-range. We surmise ionic double layers. • that with increasing concentration of the To treat the ordering effects quantita- latex particles their peripheric segment tively, Eq. [4] must be supplemented clouds are compressed and/or deformed with a structure factor S(K). For spherically when they interpenetrate each other. So the symmetric particles one finds (16): growing number of particles results in two R(K) = (1 + cos ~ 0)Y{cMP(K)S(K) [16] effects. First, the angular dependence will be less pronounced since the particles are with optically smaller. Second, the average segment density distribution between the S(K) = 1 + 47rp particles is more uniform which results in a higher average refractive index of the back× I~ r2h(r)[sin (Kr)](Kr)-ldr [17] ground and thus in a lower scattering power of the particles. These two effects and h(r) = g(r) - 1. Here g(r) is the radial complicate the extraction of S(K) from the distribution function which measures the experiments. To overcome this difficulty we probability to find a particle center in a will introduce an apparent scattering factor volume element separated at a distance r for particles in the concentrated solution from the center of a given particle. The denoted by Pc(K). This Pc(K) was obtained function h(r) is often called total correlation by drawing a smooth curve between the function. Further p = cNAM -1 is the oscillations of the measured scattering particle number density. curve, under the following restrictions: (a) Because the integral in Eq. [17] is a the Pc(K) coincides with the experimental Fourier transform, it is, in principle, possible curve for large K, because S ( K ) ~ 1 for to obtain h(r) from the reverse transform, large K; (b) Pc(K) is a straight line for small K in a semilogarithmical plot against h(r) = (2~'2p)-1 K 2, according to Guinier's approximation; × I~ Kz[S(K) - 1](Kr)-i[sin Kr]dK. [18] and (c) the resulting structure function S(K) has to satisfy the condition (23, 24) The function h(r) is determined by the interaction forces between the particles and depends further on p and T. In the next section we will consider the evaluation of S(K) and h(r) in more detail and also discuss how these functions may be related to an interaction potential with the help of some models taken from liquid state theory. a. Evaluation of S(K) and h(r) In principle, S(K) can be obtained by simply dividing the scattering results by P(K) according to Eq. [16]. Doing so, however, completely unrealistic results are obtained. It appeared that a somewhat different P(K) had to be chosen with a Journal of Colloid and Interface Science, Vol. 72, No. 2, November 1979 I f K2[S(K) - 1]dK = -2zr2p, [19] which follows from Eq. [18] by taking the limit for small r, where h(r = 0) = 1. By estimation of Pc(K) from the first two requirements (a and b), S(K) is calculated and checked by performing the integration. A trial and error procedure is used until Eq. [19] is satisfactory fulfilled. In Fig. 8 some examples are given together with the calculated values of the integral. These can be directly compared to the value of the right-hand side of Eq. [19] obtained from the concentration and the particle mass from sedimentation measurements. Clearly, STRUCTURE IN LATEX DISPERSIONS 331 .¢ Iog(R(K) /g-1 cm2) C / 3 2 - 1.( ~*'"-.. i 0 " J 2 FIG. 8. Guinier plot of the scattering sample P12 in benzene at 6.5°C (h0 = cm -z and the drawn lines are different Pc(K). Values of the integration in Eq. all in 1014 c m - L F r o m the fight-hand = R(K)/(1 + cos 2 0). - 4 , K2 //1010cm-2 6 intensity divided by concentration, R*(K)/c, versus K 2 for 546 n m ) . • are experimental data for c = 5.39 × 10-2g estimated curves for the single particle scattering function [19] are: curve 1, +0.97; curve 2, -0.92; curve 3, -1.55, side of Eq. [19] one obtains - 0 . 8 0 × 10TM cm -3. R*(K) the integral condition is a rigorous one, and some selection of the S(K) can be made. In Fig. 9 we show for P12 one calculated S(K). One of the characteristic features of this curve is the sharp peak, implying that considerable spatial ordering of the particles has occurred. Further the position of the peak, denoted by Km shifts to higher values 14 S(KJ 1.2 1.0 015 i i K/10 5 crn i 2.0 . / 08 0 . 4 ~ FIG. 9. Best fit for the structure function S(K) versus K for latex sample P12 in benzene at c = 5.39 × 10-2 g cm -3, as measured at 6.5°C and at h0 = 546 nm. Journal of Colloid and Interface Science, Vol. 72, No. 2, November 1979 332 NIEUWENHUIS AND VRIJ when the particle concentration increases, as is summarized for P10 and P12 in Table III. It appeared that the values of cKm 3 are nearly constant. This will be discussed further in the next section. In a further analysis we have determined the total correlation function, h(r), and also the so-called direct correlation function, c(r), which is defined by Fourier transformation of the structure function S(K) (23), h(r) = (27r2p)-a I f K2[S(K) - 1][sin Kr](Kr)-~dK [201 c(r) = (27r2p)-1 Io K 2 [ S ( K ) - 1]S(K)-l[sinKr](Kr-a)dK" [21] Equations [20] and [21] require knowledge of S ( K ) from K = 0 to K = ~ . Below K = 0 . 5 × 105cm - ~ ( f o r 0 = 15° a t h 0 = 546 nm) S ( K ) has to be extrapolated to K = 0. Because in all cases we found that S ( K ) equals unity for K I> 3.0 × l0 s cm -a integration is performed only up to this point. h(r) and c(r), as obtained in this way, for example P12 is shown in Fig. 10. The spurious peak in h(r) at small r is often found in these transforms [see discussion by Karnicky et al. (23)] and probably must be attributed to truncation errors or inaccuracies in the S(K). It will further be discarded. TABLE III Position of the MainDiffractionPeak in Concentrated Latex Dispersions c Km cK~ a d~ (10-2 g cm-3) (105 cm-') (10-'7 g) (nm) L a t e x P10 at 23°C (h0 = 546 nm) 7.7 1.60 5.8 1.45 4.0 1.28 2.3 1.07 1.88 1.90 1.89 1.86 480 530 600 720 L a t e x P10 at 23°C (h0 = 436 nm) 7.7 1.58 5.8 1.47 1.95 1.84 490 520 L a t e x P12 at 6.5°C (h0 = 546 nm) 5.4 1.22 3.8 1.08 2.3 0.92 3.0 3.0 2.9 630 710 840 Journal o f Colloid and Interface Science, Vol. 72, No. 2, November 1979 b. Interaction Potential The spatial ordering of the particles in the dispersion, as described in the radial distribution function g(r), is due to the interacting forces between the individual particles. The problem of relating the pair potential for interacting particles with the pair distribution function g(r), is known from the theory of simple liquids and dense gases. Starting from a given effective pair potential ~(r), Monte Carlo and Molecular Dynamical methods can be used to calculate numerically the distribution functions. Besides a number of analytical equations are derived by several authors using different types of approximations. For recent reviews see, e.g., Refs. (25) and (26). All these equations are not exact over the whole range of the particle number density as can be tested by comparison with the computer experiments. The Percus-Yevick (PY) and Hypernetted Chain (HNC) theories have been proved to work reasonably well in the investigated cases. So for the inverse route, to determine the effective pair potential ~(r) from the distribution functions we used the PY and HNC equations with/3 = (kT)-l: /3cbpy(r) = In [1 - c(r)(1 + h(r)) -a] [22] /3qbHNc(r) = h(r) - c(r) - In [1 + h(r)].[23] Equations [22] and [23] are exact in the lowdensity limit but are only approximate to STRUCTURE 333 IN LATEX DISPERSIONS 0./~ h(r) 0 , , r AO 2 nm I-0.4 -0.8 FIG. 10. T o t a l c o r r e l a t i o n f u n c t i o n h(r) for l a t e x p a r t i c l e s P12 in b e n z e n e at c = 5.39 x 10 -2 g c m -3. the behavior of denser systems. In Fig. 11 The small minimum of the order of a few we show one of the calculated pair po- tenths of k T is somewhat puzzling. In our tentials. The £b(r) found from HNC and PY systems we expect a purely repulsive pair are nearly the same. The exact shape of potential because benzene is a good solvent the resulting qb(r) must be considered with for PMMA. It seems that the application some caution, however, because of the of the PY and HNC approximations is not uncertainties along the step by step proce- reliable in this range of r. This is cordure starting from the experimental scatter- roborated by the fact that the potential of ing data. Nevertheless dp(r) shows a steep mean force, V ( r ) - - k T In g(r), which is increase below a certain interparticle center- the potential (free) energy of a pair in the to-center distance pointing to a strong presence of the other particles, is nearly repulsion between the hard particle cores at the same as the above calculated qb(r) (see r = 450 rim. Fig. 11). Because the above analysis could We note here that also some uncertainty only be accomplished with the highest can be produced by an unknown amount 2 of multiple scattering. In the Appendix a rough estimation is '\ made of the double scattering contribu- ~ ( r ) / k T I tion to the measured intensity at 90° scattering angle. When applied to our geometry the double scattering is less than 10% of the single scattering. Calculation of this effect at other angles will be much more difficult due to the asymmetric geometry of the ex\\ perimental arrangement. Moreover the influence of multiple scattering on the i I I I I I structure function will be reduced in divid4 6 8 ing the measured intensity data by the semir / 1 0 2 nm empirical curve for the apparent particle FIG. 11. E f f e c t i v e p a i r p o t e n t i a l th(r ) ( ) a n d poscattering function. So we expect only small t e n t i a l o f m e a n f o r c e V(r) = - k T In g(r) in u n i t s k T deviations in the foregoing calculations due ( - - - ) , for l a t e x p a r t i c l e s , P12 in b e n z e n e at c = 5.39 X 10 -2 g c m -3. to more than once scattered light. Journal of Colloid and Interface Science, Vol. 72, No. 2, November 1979 334 NIEUWENHUIS AND VRIJ 0.2 log ( K m a x / 1 0 5 cm-1} 0.1 - 0:1 I i ' 0.2 ~_ 016 '08 I log (c/lO-2gcm 3 ) FIG. 12. Double logarithmic plots of apparent Bragg spacing against weight concentration for the latex dispersions in benzene. Q, sample P10 at 23°C; O, sample P12 at 6.5°C. concentrations, it is also of interest and useful to discus.s the position, Km, of the maxima in S(K) with a "quasi-lattice" approach. c. Quasi-lattice Approach On the basis of a consideration of scattering data in liquid mercury, Oster and Riley (20, 27) came to the practical conclusion that for solid spherical particles the center-to-center distance of separation between neighboring particles in "expanded" liquid-type configuration is given approximately by the apparent Bragg spacing Am of the maximum in S(K). It is therefore of interest to use their analysis for our systems. The apparent spacing, Am, is derived from the angular position of the maximum by the straight forward application of Bragg's law, Am = k[2 sin (0/2)] -1 = 27rKm1. The values of Am at different stages of dilution are plotted versus the concentration (w/v) on a double logarithmic scale in Fig. 12. The linearity of the plot is a measure of Journal of Colloid and Interface Science, Vol. 72, No. 2, November 1979 the"degree of crystallinity" in the solutions due to quasi "long-range" order, while the slope = +0.33, indicating the lattice expansion on dilution is in three dimensions. The relation between the Bragg spacing, Am, of the principal peak and the mean interparticle distance, do, is given for a closest cubic packing (fcc) by (20): de = (3/2)1/2Am. [24] Numbers for dc are shown in Table III. From the number of lattice points per unit cell (=4), M follows from M = (3/4) X 31/2NACA3m,and was found to be 3.7 x l09 g mole -1 for sample Pl0 and 5.8 × 109g mole -1 for sample P12. Both values are in reasonable accordance with those obtained from sedimentation plus diffusion. d. Effective Diffusion Coefficient Structure formation not only shows up in the light scattering intensity as a function of 0 but in the quasi-elastic light scattering as well. In Fig. 13, it is shown how Deff as a function of K is also structured at higher S T R U C T U R E IN L A T E X D I S P E R S I O N S 3.0 1.5 D o / D e f f (K) S(K) 1.0 335 2.0 1 ----,-.-- 0.5 K/lO cr( 1.0 FIG. 13. Structure function S(K) and relative reciprocal effective diffusion coefficient Do/Dell(K) (Q) v e r s u s K , for P10 latex particles in b e n z e n e at c = 7.7 × 10-2 g c m 3. Do is the " i n d e p e n d e n t particle" diffusion coefficient for c --~ 0. concentrations. Recent theoretical (11) and experimental (21, 22) work has shown that S ( K ) can be inferred indirectly through the dynamics of concentration fluctuations. In Fig. 13 one observes that the extrema in D~-f} coincide with those of S(K). This result is in good agreement with those of others (21, 22) for latex dispersions in extremely dilute electrolyte solutions. We think that in both cases the interaction must be characterized as soft, long-range repulsions. The discrepancy between the static structure factor and D~-f}in our systems has to be attributed to an additional term to the instantaneous velocity of a particle in an interacting dispersion: the hydrodynamic interaction. The theories on this point are not yet worked out in detail. 7. C R Y S T A L S In this section we will describe the crystal-like structure which we observed in a test tube with a dispersion of sample P12 after standing for several months. When a moderately concentrated latex dispersion was set apart, the particles settled to the bottom of the container by gravity, leaving an almost clear solution in the upper part of the tube. The sediment consisted of ordered arrays of particles. This crystal- like structure occupied a relative large volume and no more changes take place after a few months. Under more detailed investigation the ordered sediment appeared to be very transparent and, when placed in a beam of white light, very strong and sharp diffraction colors could be observed at different angles (see Fig. 14). Between the "particle-free" supernatant and the concentrated crystal-like sediment we observed a small region that showed the behavior of a highly concentrated (liquid-like) dispersion. To determine the distances between the net planes (d), we placed the sample in a beam of monochromatic light and measured the angle (0) of the first diffraction peak. According to the Bragg relation for the first order diffraction, 2d sin (0/2) = X, the distance d can be calculated from 0, and the wavelength of the light in the medium X (=X0/n). In Table IV the results are given for three different wavelengths. Assuming an fcc lattice the observed diffraction corresponds to the reflection from the (111) plane; reflections from (100) and (110) are destroyed by geometric structure effects. In this case the center to center distance, dcr, of the lattice points can be calculated from the distance between the net planes: dcr = (3/2)~/2d, and becomes d e r = 418 nm. Journal of ColloM and Interface Science, Vol. 72, No. 2, November 1979 336 NIEUWENHUIS AND VRIJ FIG. 14. Diffraction of white light by a crystal-like latex sediment of sample Pl2. The test tube is placed in flask with toluene to avoid reflection of light by the glass wall. The incident light beam passes through the test tube from right to left. The transmitted beam gives an oval white spot on the glass wall of the flask. A: blue; B: green; C: yellow; D: red. TABLE IV Crystal Diffraction Data for Latex Sediment of Sample P12a 8. FINAL DISCUSSION h0 (rim) n sin 0 d (nm) 436 546 633 1.521 1.503 1.495 0.418 0.535 0.621 343 339 341 h0, Wavelength of incident light in vacuo; n, refractive index of benzene; 0, angle of diffraction; and d distance of net planes. Journal of Colloid and Interface Science, Vol. 72, No. 2, November 1979 A n o v e r a l l v i e w o f the r e s u l t s o b t a i n e d in the p r e v i o u s s e c t i o n s will b e g i v e n in t e r m s o f the d i f f e r e n t sizes a n d d i s t a n c e s o f i n t e r a c t i o n o f the p a r t i c l e s as s u m m a r i z e d in Fig. 15. W h e r e a s a d e t a i l e d p i c t u r e o f the i n t e r n a l p a r t i c l e s t r u c t u r e is n o t k n o w n , all the m e a s u r e d p a r t i c l e sizes are e x p r e s s e d as diameters of apparent, homogeneous spheres (i.e., G u i n i e r light s c a t t e r i n g , q u a s i - e l a s t i c STRUCTURE IN LATEX DISPERSIONS P12 rim 9oo - 800 P10 nm c/gcm-3 %1 0.023 ~ 0.09 Ool/dc 700 . 337 0.20 c/gcm-3 ~ ~I 0.023~ 008 !!!!yd c o,lz. 5oo - 0,20 VISCOSITY 500 DIFFUSION [C6H5} ~/kT: 1 - 400 - - CRYSTAL GUiNIER / 3oc 0.27 - - VISCOSITY - - - DIFFUSION (C6H6} OUINIER PMMA DIFFUSION (H201 OUINIER OVB - - DIFFUSION (H20} EM 2oc - - EM lOC Fro. 15. Diameters and distances in PMMA dispersions. EM: 2aEM(H20); Guinier DVB: 2(5/3)l/2rgA; diffusion (H20): 2aD(H20); Guinier PMMA: 2(5/3)lmrgB; crystal: dcr; q~ = lkT from Fig. 11; diffusion (C6H6):2aD = kT(37r'oD)-~; viscosity: 2an; d c = 27r(3/2)11ZKml, all in nm. light scattering, intrinsic viscosity). From the numbers found in Fig. 15 it seems reasonable to conclude that the particles have a rather open crosslinked kernel with a much less dense periphery. The periphery does not contribute much to the light scattering but still strongly obstructs the motion of solvent. At higher particle conc e n t r a t i o n s - e v e n near the closest packing volume fraction: cbn = (4/3)~-a~--the dispersions still showed a relatively low viscosity as compared with high-molecular-weight polymer solutions of the same weight concentration from which we conclude that the particles still behave as separate entities and that interchain entanglements between overlapping peripheries are absent. From Figs. 10 and 11 we conclude that interparticle repulsions of - lkT arise when the interparticle distance reaches 450 nm (for P12), i.e., somewhat below the hydrodynamic diameter (2an and 2a9), which seems acceptable. The crystal diffraction data of P12 show an even smaller inter° particle distance of 420 nm (with ~b -1.5kT from Fig. 11). This implies a marked overlap of the peripheries after settling of the dispersion in the terrestrial gravitational field. Let us turn the attention now to the spatial ordering of the concentrated (liquid) dispersions. The relevant distances de as obtained from the quasi-lattice approach (see Section 6) are given in Table III and also in Fig. 15. Here (see [24]) dc = (3/2)1/2Am = 7.695 Km1, and Kin is the position of the maximum in S(K). One may doubt whether dc has any significance beyond the concept of a "lattice." It can be argued, however, that d~ must be very close to the position, rm, of the principal maximum in the radial distribution function g(r). For many liquids--rather independent of density--it is found (empirically) that Kmr m = 7 to 8, e.g., liquid argon (28): 7.1; dilute Journal of Colloid and Interface Science, Vol. 72, No. 2, November 1979 338 N I E U W E N H U I S A N D V R IJ aqueous latex (29): 7.4; our system (see Figs. 9 and 10) = 7.6; computer simulation with a purely repulsive pair potential (30): 7.40; the so-called Ehrenfest rule (the position of the minimum in (Kr) -1 sin (Kr)): 7.73. The large variation of dc is remarkable, for both P10 and P12, with concentration (see Figs. 12 and 15), cd~ being approximately constant and with very large values of de, particularly at the smaller concentrations. This implies that ordering takes place at distances much larger than the hydrodynamic diameters. Both the proportionality of d~ 3 with c and the very large values of de were also reported for charge-stabilized latex in water (29, 30). This was explained by the very extended electrical double layers in these systems, which contained a very small amount of electrolyte. This suggests that ordering by electrical charge effects also occurs in our systems. Others (31, 32) have described the occurrence of charge effects in nonpolar media especially in relation to inorganic colloids. It is unquestionable that these effects, when present, give rise to forces of a very long range (say - 1 / ~ m ) , because the number of charges can only be very small in solvents with a low dielectric constant. We investigated the presence of free charges with some simple electrical conductivity measurements. Conductivities down to about 10-14 12-1 cm -1 were measured with an accuracy of about 5% by a simple dc method using a Jones-type cell with a cell constant of 16 cm (33, 34), a Fluke 412 B power supply, and a Keithley 602 electrometer. For benzene (Baker Analyzed) without further purification we found a specific conductivity ~Cs~c~-8 × 10-14 12-1 cm -1 and for P12 latex dispersions Kso~c= 16 × 10-14 and 30 x 10-14 12-1 cm -1 for c = 0.006 and 0.010 g cm -3, respectively. Whereas nothing is known about the distribution of charges over particles and counterions let us assume that all the charges are present on the latex Journal of Colloid and Interface Science, Vol. 72, No. 2, November 1979 particles alone, which will give an overestimation of the number of charges per particle. Using the equation Ks~c = F2z2D × ( R T ) - l c M -a (see e.g., (35)), where z is the average number of elementary charges per particle and F is Faraday constant one finds from the values of D and M determined above: z2 1011Kspeec-1. This gives a z value of the order of unity. The presence of (much smaller) counterions would even give a much smaller number. We assess that, e.g., the well in V(r) given in Fig. 11 would require a z value of 5 or 10. As a consequence we conclude that free electrical charges cannot be important in our systems and that electrostatic repulsion seems no plausible explanation for the longrange ordering in the dispersions. Further research is required to elucidate the nature of the interactions at intermediate concentrations. In the meantime we have started static and dynamic light scattering experiments with somewhat smaller microgel particles (36). = APPENDIX: E V A L U A T I O N OF DOUBLE SCATTERING Because no appropriate treatment was found in the literature for the description of multiple scattering in our experimental geometry, we will make a rough estimate of the contribution of the double scattered intensity. The experimental set-up is schematically represented in Fig. 16. Calculations are made only for the scattering angle 90° . In this arrangement, single scattering originates only from the scattering volume in the center of the cell. Double scattered radiation can reach the photomultiplier from a second scatterer situated in the detection path of the photomultiplier. The incident light on this scatterer comes only from a first scatterer somewhere in the incident beam. The dimensions of the incident beam and the detected beam are taken equal. In the figure, dl represents the width of the beams in the plane of the STRUCTURE IN LATEX DISPERSIONS 339 ? ?2 / \ I dl t FIG. 16. Experimental setup in the light scattering apparatus. The picture is given for an observation angle of 90°. The Roman numerals refer to the four quadrants. Also given is a possible path for double scattered light. For explanation of symbols see text. figure and d2 the height in the perpendicular direction. The first scattered intensity at an angle 02, with the direction of the incident beam by a volume element d~dsdx in the incident beam at a distance r from the volume element, can be expressed as: dI1 = Io•r-Sp X e x p [ - h s sins (Oi/2)]dxdsdx. [A1] Here Io is the intensity of the incident beam, Y[ is an optical constant (see Eq. [5]), p is the particle number density, and the exponential term is the Gulnier approximation for the particle scattering function, with h 2 = (4zrna~t-1)s/5 (see Eq. [13]). If this light is scattered again by particles in the detection path the intensity is given as d[2 = dI~ Y[R-Sp × e x p [ - h s sins (02/2)]d~d2dy. the first and second scattered radiation. Note that we have neglected the pathlength of the light in the cell with respect to the distance R. Let us first determine the contribution of the scattering center to the double scattering. With the help of calculations performed by Blech (37) for cylindrical samples it is easily shown that the secondary scattering in the scattering volume is less than 2% of the single scattering when applied to our data. So we will calculate double scattering from the parts of the beams outside the center. By substituting Eq. [A1] in Eq. [A2], and some algebra, the following relation is obtained dis = Io( ~{d~d2p)2R-2r -s exp(-h s) × exp[hS(cos 01 + cos 02)/2]dxdy [A2] Here R is the distance of the scattering cell to the detector, and 02 is the angle between [A3] which must be integrated over both axes. Figure 16 is divided into its four quadrants. Because of the angular dependency of the Journal of Colloid and Interface Science, Vol. 72, No. 2, November 1979 340 NIEUWENHUIS AND VRIJ scattering of the particles the contribution of the double scattering across each quadrant is different. By using simple geometrical properties and changing from rectangular to polar coordinates one obtains the following equation, where d12 2/2 is the smallest distance between incident and detection beam and 122/2 is the largest distance in the cell. 12 --= Io( ff~dld2p)2R -2 × e x p ( - h 2) In (21d~2)Q, [A4] where Q = Q I "-[- QII + QIII + Qlv, as contributions from each of the quadrants. In our case only Qin is of importance, ACKNOWLEDGMENTS The authors wish to acknowledge Dr. W. van der Drift for his advice and assistance in preparing the latex particles. We also thank Mr. J. Pieters of the Center of Electronmicroscopy of the Department of Biology for making the electron micrographs. We are indebted to Dr. C. Pathmarnanoharan for helpful discussions on multiple scattering. Grateful thanks are extended to Professor J. Th. G. Overbeek for helpful suggestions on the electric conductivity measurements. Dr. H. Fijnaut and Mr. H. Mos are gratefully thanked for making the diffusion experiments and Mr. J. Suurmond for making the ultracentrifuge measurements. Finally we thank Miss H. Miltenburg and Mrs. M. uit de Bulten for typing the manuscript and Mr. W. den Hartog for drawing the illustrations. REFERENCES Qm = If exp[h2(cos 01 + sin 00/2]d01. [A51 Because no analytical solution is known for this type of integrals, and the sum cos 01 + sin 02 changes only slightly in the interval of 0, we took an average value for this sum of 1.2. Then one obtains QIII -~ (~r/2) exp(3h2/5). [A6] For our samples where h 2 ~ 12 the sum of the other Q's is less than 10%. The total single scattering by the center of the cell is given by 11 = IoY{d~d2R-2p exp[-h2/2] [A7] and the ratio of double to single scattering becomes 12/12 = (~r/2)Rod2 In (21/dl) x exp(h2/10) [A8] where Ro = 3fp is the Rayleigh ratio at zero scattering angle. For our geometries l = 1 era; d2 = 0.16 cm one finds 12/12 = 2.1 × (R0/cm-2). In our experiments R0 = 0.03 cm -1 for the highest concentrations (see Fig. 7). So we conclude that the multiple scattering is, say, less than 10% of the total scattering intensity (at 90 ° scattering angle). Journal of Colloid and Interface Science, Vol. 72, No. 2, November 1979 1. Shashoua, V., and Beaman, R., J. Polymer Sci. 33, 101 (1958). 2. Hiltner, P. A., and Krieger, I. M., d. Phys. Chem. 73, 2386 (1969). 3. Kose, A., and Hachisu, S., J. Colloid Interface Sei. 46, 460 (1974). 4. Hansen, J. P., and McDonald, I. R., "Theory of Simple Liquids." Academic Press, London, 1976. 5. Vrij, A., Pure Appl. Chem. 48, 471 (1976). 6. Vrij, A., Nieuwenhuis, E. A., Fijnaut, H. M., and Agterof, W. G. M., Disc. Faraday Soc. 65, 101 (1978). 7. Ono, H., and Saeki, H., Colloid Polym. Sci. 253, 744 (1975). 8. Hul, H. Van Den, and Vanderhoff, J., J. Colloid Interface Sci. 28, 336 (1968). 9. Fitch, R., and Tsai, C., in "Polymer Colloids" (R. Fitch, Ed.), Vol. III. Plenum Press, New York, 1971. 10. McDonald, S., Daniels, C., and Davidson, J., J. Colloid Interface Sci. 59, 342 (1977). 11. Berne, B. J., and Pecora, R., "Dynamic Light Scattering." Wiley, New York, 1976. 12. Tanford, C., "Physical Chemistry of Macromolecules," Chap. 22. Wiley, New York, 1963. 13. Schulz, G. V., and Hoffmann, M., Makromol. Chem. 23, 220 (1957). 14. Guinier, A., and Fournet, G., "Small Angle Scattering of X-rays," p. 126. Wiley, New York, 1955. 15. Kerker, M., "The Scattering of Light and Other Electromagnetic Radiation," p. 427. Academic Press, New York, 1969. 16. Riley, D. P., and Oster, G., Disc. Faraday Soc. 11, 107 (1951). STRUCTURE IN LATEX DISPERSIONS 17. Benoit, H., and Froelich, D., in "Light Scattering from Polymer Solutions" (M. B. Huglin, Ed.), Chap. 11. Academic Press, London, 1972. 18. Cordier, P., J. Chim. Phys. 64, 133 (1967). 19. Brandrup, J., and Immergut, E. H., "Polymer Handbook," 2nd ed. Wiley, New York, 1975. 20. Oster, G., Receuil trav chim. 68, 1123 (1949). 21. Brown, J., Pusey, P., Goodwin, J., and Ottewill, R., J. Phys. A: Math. Gen. 8, 664 (1975). 22. Schaefer, D., J. Chem. Phys. 66, 3980 (1977). 23. Karnicky, J., Hollis Reamer, H., and Pings, C., J. Chem. Phys. 64, 4592 (1976). 24. Krogh-Moe, J., Acta Crystallogr. 9, 951 (1956). 25. Barker, J. A., and Henderson, D., Rev. Mod. Phys. 48, 587 (1976). 26. Andersen, H. C., Chandler, D., and Weeks, J. D., Adv. Chem. Phys. 34, 105 (1976). 27. Oster, G., and Riley, D., Acta Crystallogr. 5, 1 (1952). 28. Mikolaj, P., and Pings, C., J. Chem. Phys. 46, 1412 (1967). 29. Brown, J., Goodwin, J., Ottewill, R., and Pusey, 30. 31. 32. 33. 34. 35. 36. 37. 341 P., "Colloid Interface Sci.," Vol. IV (Hydrosols and Rheology). Academic Press, New York, 1976. Megen, W. Van, and Snook, J., J. Chem. Phys. 66, 813 (1977). Osmond, D. W. J., and Waite, F. A., "Dispersion Polymerization in Organic Media" (K. E. J. Barrett, Ed.), Chap. 2. Wiley, London, 1975. Lyklema, J., Advan. Colloid Interface Sci. 2, 65 (1968). Koelmans, H., and Overbeek, J. Th. G., Disc. Faraday Soc. 18, 52 (1954). Albers, W., and Overbeek, J. Th. G., J. Colloid Sci. 14, 501 (1959). Macinnes, D. A., "The Principles of Electrochemistry," Chap. 3. Dover Publications, New York, 1961. Fijnaut, H. M., Pathmamanoharan, C., Nieuwenhuis, E. A., and Vrij, A., Chem. Phys. Lett. 59, 351 (1978). Blech, I. A., and Averbach, B. L., Phys. Rev. 137, A 1113 (1965). Journal of Colloid and Interface Science, Vol. 72, No. 2, November 1979