Pigment

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 10

Colloid & Polymer Sci. 254, 726-735 (1976) 1976 Dr. Dietrich Steinkopff Verlag GmbH & Co.

. KG, Darmstadt ISSN 0303-402X / ASTM-Coden: CPMSB (formerly KZZPAF)

Industrial ResearchLaboratories, Kao Soap Company, Wakayama-shi (Japan)

Stabilities of aqueous inorganic pigment suspensions


N. M o r i y a m a

With 9 figures and 4 tables (Received July 25, 1974)

Introduction
Although much studies have been reported upon the stabilities of aqueous pigment suspensions in the presence of surfactants (1-9), systematic studies upon the stabilities in view of the chemical structure and the concentration of surfactants, zeta potential of pigment particles etc. have been scarcely found (10). In the preceding paper (11), a detailed investigation on the stabilities of ferric oxide suspensions in solutions of various types of surfactants and inorganic phosphates, and its correlation with the zeta potential of ferric oxide particles, the adsorption of the surfactants or the inorganic phosphates onto the particles, and pH of the suspensions have been reported. The present paper describes the effects of various types of anionic surfactants and inorganic phosphates on the stabilities of aqueous suspensions of titanium dioxide and others, and the discussion upon the stabilities from the view-point of surface chemistry such as zeta potential and adsorption. It was further attempted to interpret the stability data in relation with D, L.V. O theory.

Experimental
Materials. Sodium salt of polyacrylic acid (PA) was prepared according to the method of Kawamura et al. (12). Sodium salt of styrenemaleic acid copolymer (SM) was prepared by polymerization of styrene and maleic anhydride in toluene using benzoyl peroxide as initiator and dodecyl mercaptane as polymerization conditioner, and by hydrolysis with sodium
T 178

hydroxide. Sodium salt of ~-olefine (Cs)maMc acid copolymer (OM) was prepared by polymerization of a-olefine (C8) and maleic anhydride in toluene using benzoyl peroxide as initiator, and by hydrolysis with sodium hydroxide. Sodium salt of formalin condensate of fi-naphthalene sulfonic acid (NSF) and sodium salt of formalin condensate of alkyl (C4) naphthalene sulfonic acid (A1-NSF) were prepared according to the method of Hattori et al. (13), respectively. The average molecular weights of vinyl copolymers and formalin condensates described above were calculated from vapor-pressure depression data. The vapor-pressure depression measurements were made with an Osmometer, Hitachi-Perkin Elmer Model 115, using methyl ethyl ketone as solvent for SM, OM, NSF and A1-NSF, and chloroform as solvent for PA. The vaporpressure measurements for SM and OM were made using the samples before hydrolysis of them which were purified by pouring the small amount of methyl ethyl ketone solutions of the samples into the large amount of toluene. The measurement for PA was made using the sample obtained by methylation of PA after its ion-exchange reaction from Na+-form to H+-form. Those for NSF and A1-NSF were made using the samples obtained by sulfochlorination of them with thionyl chloride according to the method of Zol[inger (14). The average molecular weights thus obtained are listed in tab. 1. Sodium alkyl (C,2) benzene sulfonate (LAS) and sodium alkyl (C4)-naphthalene sulfonate (ANS) were prepared according to the ordinary method (15). Sodium dodeeyl sulfate (SDS) was prepared and purified according to the procedure described earlier (16). Ortho-

Moriyama, Stabilities of aqueous inorganicfligment suspensions


Tab. 1. Chemical composition, average molecular weight and abbreviation of vinyl copolymers and formalin condensates Type Chemical composition Average Abbremolecularviation weight PA SM

727

Vinyl Na-salt of 3100 copolymers polyacrylicacid Na-salt of 4500 copolymer of styrene and maleic acid Na-salt of 4100 copolymer of ~-olefine (C8) and maleic acid Formalin Na-salt of 900 condensates formalin condensate of fl-naphthalene sulfonic acid Na-salt of 1100 formalin condensate of alkyl (C4) naphthalene sulfonic acid

OM

NSF

A1-NSF

phosphate (NaH2PO4), pyrophosphate (Na2H2P207 and Na4P207) and tripolyphosphate (NasPaO10) were of reagent grade. Above all phosphates were obtained from Wako Pure Chemicals Co. Pigments. Titanium dioxide, zinc oxide, chrome yellow, aluminium hydroxide and clay were used as inorganic pigments. Various physical properties of them are listed in tab. 2. All pigments listed in tab. 2 were washed with distilled water until there was no change

in the conductivity of the filtrate. Specific surface areas of them were determined by the B.E.T. method using nitrogen at - 1 9 5 C . Isoelectric points of the pigments were determined according to the directions of Parks et al. (17). Average diameter of the primary pigment particles were calculated fromelectron microscopic observation. Sedimentation velocity measurements. Aqueous solutions of the materials, which refer to surfactants and inorganic phosphates, were prepared by dissolving the desired quantities of the materials in distilled water of 250 ml. About 30 ml of the material solutions thus prepared was poured into graduated test cylinders of 30 ml capacity and 30 cm height, and maintained in a thermostat at 25 C. Into the above cylinders, 0.30 g of the pigment was added. After the cylinders were turned endover-end several times, the suspensions were calibrated to 30 ml by adding the material solutions. The cylinders were shaken 50 times vigorously and allowed to stand for 24 hours to attain the equilibrium. They were also shaken 50 times vigorously before the measurements. The positions of the sedimentation boundary between transparent solution parts or very dilute suspension parts and the suspension parts, formed by the sedimentation of the pigment particles, were measured periodically. 171eclrophoretic mobility measurements. Mobility measurements of pigment particles were made using a Zeta-Meter (Zeta-Meter Inc.) by diluting a quantity of the same suspensions used for sedimentation velocity measurements into the supernatant solutions of their sus-

Tab. 2. Various physical properties of inorganic pigments Pigment titanium dioxide zinc oxide chrome yellow aluminium hydroxide clay Specific surface areas (m2/g) 5.5 4.0 3.5 3.2 1.2 Average particle size (/z) 0.4 0.3 0.5 0.6 0.5 Isoelectric point 6.0 8.4 9.7 -Maker Ishihara Industries Co. (Tipaque R-550) Hakusui Chemicals Co. (Special grade) Kikusui Dyes Co.

(ODG)
Showa Denko KK (Higilite H-42) Nozaki ClayCo. (ssw)
48

728

Colloid and Polymer Science, Vol. 254 No. 8 (1976) (cm


3O I i I i i

pensions obtained by removing the particles with refrigerated-automatic Ultracentrifuge, Serval Model RC2-B. Adsorption measurements. The amounts of the adsorption of the materials onto the pigment particles w e r e calculated from measurements of the difference in concentrations of the material solutions before and after preparing the suspensions. The material solutions after preparing the suspensions above referred to are the same as those used for electrophoretic mobility measurements. T h e concentrations of the surfactants having benzene ring or naphthalene ring in the molecule, among the material solutions, were calculated by measuring the intensity of absorption band at about 260 m#, (256 m# for LAS, 288 m/, for NSF, etc.) using Shimadzu Spectrophotometer UV200. On the other hand, those having no benzene ring or no naphthalene ring in the molecule were measured with TOC Meter, Toshiba-Be&man Model 102, which determined the amount of carbon dioxide generated by firing the solutions at 950 C, using CoO as catalyst. The concentrations of orthophosphate and polyphosphate solutions were measured according to the direction of 2Vakahara (5). pI-I measurements, pH of the same suspensions used sedimentation velocity measurements were measured with pH Meter, TO A Electronics Ltd. Model HM-7A. pH of the suspensions were controlled by adding the aqueous sodium hydroxide solution or aqueous hydrochloric acid solution.
Results and discussion

|
E i
E 10 ~.003 20

I
0.O7

~1
lO standing time, t(hr)

I
50

I
100

Fig. 1. The position h of sedimentation boundary of the suspension vs. standing time t for various concentrations of SDS at 25 C. The numbers along the curves refer to the mole concentrations of SDS

the increase of concentration and reaches almost constant value. With further increase in concentration, it decreases again remarkably. In other words, two inflection points appear in the log TII~, vs. log C curves. One inflection point at low concentration is tentatively designated as Cl and an other one at high concentration as C1. These two inflection points as well as the absolute values of T1/2 are closely related with the stability of the suspensions. In the concentration ranges from C, to C1 the suspensions become most stable, whereas in the concentration ranges below C~ or above C1 they become unstable. In tab. 3, C/ and

Fig. 1 shows the plots of the position, h, of the sedimentation boundary of aqueous titanium dioxide suspensions vs. standing time, t, for various concentrations of SDS. The h is lowered with the t. The velocity of its lowering depends upon the t and the concentration of SDS. The time in which h falls from 30 cm height to 15 cm height is calculated from h vs. t curves and designated as sedimenting time, T1/~. In the present paper, the sedimenting time has been used as a measure of the stability of the suspensions. Fig. 2 shows the relation between the T1/2 vs. the logarithms of the concentrations for SDS, PA and NSF. In the regions of low concentrations, the sedimenting time increases with

0.0001

0.001

0.01

0.10

log. C, (M/L)

Fig. 2. Sedimenting time Tll2 vs. log concentrations C for SDS (C)), PA (@) and N S F ((]D) at 25 C. The concentrations of P A and N S F refer to the mole concentrations expressed per monomer unit of them

Moriyama, Stabilities of aqueous inorgctnicpigment suspension, Tab. 3. C~ and C1 obtained from log Tll2 vs. log C curves for titanium dioxide/water suspensions Material vinyl copolymers PA SM OM formalin condensates NSF A1-NSF general surfactants SDS LAS ANS orthophosphate NaH2PO4 polyphosphates Na2H2P207 Na4P~O7 Na5PsOlo C; (M/L) 0.00050 0.00040 0.00040 0.00032 0.00030 0.0060 0.0015 0.017 0.00042 0.000050 0.000040 0.000025 C~ (M/L) 0.021 0.012 0.012
I

729

1000 IO0 ~ 10 1
60

40
2O

0.21 0.20 0.055 0.075 0.034

~
~

20
lO

I1.

0.0021 0.014 0.012 0.0054

I0 9 8 7 6

0.0001

0.001

0.01

0.10

log concentration of surfactant, C (M/L)

*) Concentrations (M/L) of vinyl copolymers and formalin condensates are expressed in those of monomer unit of them, respectively.

C1 obtained from log T,/2 vs. log C curves for the various types of surfactants and inorganic phosphates are listed. The curves show generally distinct inflection points at both Ci and C1. It can be seen from tab. 3 that large differences in concentrations between Ci and C1 exist in the curves for the surfactants having molecular weight more than several hundreds, e.g., PA, SM, OM and NSF, and also for polyphosphates, whereas little differences between C~ and C1 exist in the curves for surfactants such as SDS, LAS and ANS, and for orthophosphate. With the surfactants of the type of vinyl copolymers, the values of C~ tend to be lowered with the increase in their hydrophobic properties. With the surfactants of the type of formalin condensate, C~ shows somewhat smaller values compared with those of the surfactants of the type of vinyl copolymers, whereas C1 shows much larger values compared with those of the surfactants of the type of vinyl copolymers. With polyphosphates, C~ shows smaller values compared with all the surfactants described above, while C1 shows almost equal values with those of the surfactants of the type of vinyl copolymers. With polyphosphates, both C~ and C1 tend

Fig. 3. Sedimenting time T~I2, zeta potential, amount of adsorption, and pH in aqueous titanium dioxide suspensions vs. log concentration for SDS (C)), PA (@) and NSF ((]p) at 25 C. The concentrations of PA and NSF refer to the mole concentrations expressed per monomer unit of them. A dotted and a broken line refer to the T~I2 vs. log concentration curves for SDS at a controlled pH of 7.7 and for PA at a controlled pH of 9.3, respecnvely to be lowered with the number of condensations. In fig. 3, the Tl/2, the zeta potential, the amount of adsorption, and pH in titanium dioxide suspensions are plotted as a function of the concentration of SDS, PA and NSF, respectively. The calculation of zeta potential, ~, from electrophoretic mobility, u, measurements is simple only when na is very large (>> 100) or very small (40.1) (18), where n is the reciprocal of the thickness of double layer and a the particle radius. As we treat nearly spherical pigment particles (from microscopic observation) and the systems of a a > 2 0 0 , tge/mholtz-Smoluchowski's equation can be employed in the form: Ege
u [1]

4:o1 where ~/and e are the viscosity and dielectric constant of the medium, respectively, and E is the applied electric field. As seen in fig. 3, the values of T1/2 of titanium dioxide particles in solutions of very
48*

730

Colloid and Polymer Science, Vol. 254 No. 8 (1976)

low concentration of SDS are much smaller than those of PA and NSF, that is, titanium dioxide suspensions in aqueous solutions of SDS are very unstable in comparison with those of PA and NSF. In the region of low concentrations, the absolute values of zeta potentials of the titanium dioxide particles in aqueous solutions of SDS, PA and NSF become large as the values of Tt/2 become large. In this region of low concentration, a correlation can be found between stability data and zeta potential data or adsorption data. In the region of high concentrations, the value of T1/~ decreases remarkably with the increase in concentration, that is, the suspension becomes unstable with the increase in concentration, whereas the zeta potential remains almost constant or slightly decreases and the amount of adsorption remains increasing. In this region of high concentrations, any correlations between stability data and zeta potential data or adsorption data cannot be found. Marked decrease in stability at high concentration may be attributed to the decrease of electrical repulsion owing to the compression of the thickness of the double layer. The thickness of the double layer, i.e., l/n, is inversely proportional to the square of the valency of the material ions and to the concentration of the material ions (18). According to the D . L . V . O . theory (19), the electrical repulsion which disturbs the coagulation of pigment particles in the disperse systems depends upon the absolute value of zeta potential and the thickness of the double layer, as described later. As seen in fig. 3, the zeta potentials and the concentrations becoming stable suspensions are different with the sorts of the surfactants. These may be explained by the above facts that electrical repulsion, i.e., stability, depends upon the thickness of the double layer and the absolute values of zeta potential. Fig. 3 shows that pH values of the suspensions dispersed.with each surfactant of SDS, PA and NSF undergo very little changes with increase in concentration of them, except those in the region of low concentrations. A dotted and a broken line shown in fig. 3 refer to the Tlt~ vs. concentration curves for SDS and PA at controlled pH (7.7 for SDS and 9.3 for PA), respectively. A dotted and a

1000,
. ~ lOO ~ - ~ lO
1 _ 60

~g

20

8~

lO

0.0001

0.001

0.01

0.10

log concentration of surfactant, C(M/L)

Fig. 4. Sedimenting time TaI~, zeta potential, amount of adsorption, and pH in aqueous titanium dioxide suspensions vs. log concentration for A1-NSF ((])), ANS (ID) and LAS (Q) at 25 C. The concentrations of A1-NSF refer to the mole concentrations expressed per monomer unit. A dotted line refers to the Tll2 vs. log concentration curves for LAS at a controlled pH of 7.8 a broken line are in accord with each solid line of SDS and PA which are not p H controlled. These results show that stabilities of the suspensions dispersed with the above three surfactants undergo only very little effects of pH. In fig. 4, the T1/2, the zeta potential, the amount of adsorption, and p H in titanium dioxide suspensions are plotted as a function of the concentrations of A1-NSF, ANS and L A S , respectively. A1-NSF is a formalin condensate of ANS. The number of condensation of A1-NSF calculated from average molecular weight shown in tab. 1 is 3.7. The value of Ci for A1-NSF is much lower than that for ANS, and the difference in concentration between C~ and C1 for A1-NSF is much larger than that for ANS, as seen in fig. 4. The absolute value of zeta potential and the amount of adsorption for A1-NSF are larger than those for ANS over all concentration ranges tested. At the same concentrations, p H of each aqueous solutions of A1-NSF and ANS containing 1 wt.% titanium dioxide show almost equal values. These results show that the adsorbability of

Moriyama, StabzTities of aqueous inorganicpigmen/ suspensions

731

the surfactant onto the pigment particles increases with increase in molecular weight and stability is highly dependent upon the adsorbability of surfactant and its contribution to zeta potential. With LAS, stability, zeta potential and adsorbability show about intermediate behaviors between those of A1-NSF and those of ANS in the region of low concentrations. The differences of the above behaviors between LAS and ANS may be due to the stronger adsorbability of LAS onto the pigment compared with that of ANS. A dotted line shown in fig. 4 refers to the T1/2 vs. log concentration curve for LAS at a controlled pH of 7.8. A dotted line is in accord with solid line of LAS shown in fig. 4. This result and fig. 4 show that the stabilities of the suspensions dispersed with the above three sudactants undergo only little effects of pH. In fig. 5, T1/2, the zeta potential, the amount of adsorption, and pH in titanium dioxide suspensions are plotted as a function of the concentrations of sodium dihydrogen phosphate (NaH2PO4), acid sodium pyrophosphate

,ooo

6o
m

20

II

II

il

.~

20

6
4

o.oool

o.ool

O.Ol

log concentPation of surfaetant, C(MIL)

Fig. 5. Sedimenting time Tll2, zeta potential, amount of adsorption, and pH in aqueous titanium dioxide suspensions vs. log concentrations for acid sodium pyrophosphate (Na2H2P207 ~ ) , sodium pyrophosphate (Na4P=Ov ~ ) , and sodium dihydrogen phosphate (NaH2PO4 (~)) at 25 C. A dotted and a broken line refer to the Tll2 vs. log concentration curves for acid sodium pyrophosphate at a controlled pH of 4.5 and for sodium pyrophosphate at a controlled pH of 9.8, respectively,

(Na2H2P~Ov) and sodium pyrophosphate (Na4P2Ov), respectively. Acid sodium pyrophosphate and sodium pyrophosphate show relatively similar behavior for T1/2, zeta potential and adsorption, while sodium dihydrogen phosphate shows much different behavior. As seen in fig. 5, the values of pH of inorganic phosphate solutions containing 1 wt % titanium dioxide particles change with the increase in their concentrations. At the same concentrations, pH of aqueous solutions of sodium dihydrogen phosphate and acid sodium pyrophosphate containing 1 wt% titanium dioxide particles show almost equal values. On the other hand, pH of sodium pyrophosphate solutions show considerably high values in comparison with those of above two phosphates. The pH values of the suspensions dispersed with sodium dihydrogen phosphate or acid sodium pyrophosphate are lower than the isoelectric point (pH 6.0) of titanium dioxide particles. However, titanium dioxide particles have negative charge, as shown in fig. 5. This may be due to the adsorption of pyrophosphate ion or phosphate ion having negative charge. A dotted and a broken line shown in fig. 5 refer to the T1/2 vs, concentration curves for acid pyrophosphate and sodium pyrophosphate at controlled pH (4.5 for acid sodium pyrophosphate and 9.8 for sodium pyrophosphate), respectively. With acid sodium pyrophosphate, the suspensions at not controlled pH (changing from 7.0 to 4.5), i.e., solid lines shown in fig. 5 are more stable than those at a controlled pH of 4.5, i.e., dotted lines shown in fig. 5. With sodium pyrophosphate, the suspensions at a controlled pH of 9.8, i.e., shown in fig. 5, are more stable than those at not controlled pH (changing from 7.0 to 10.0), i.e., solid lines shown in fig. 5. These results show that the stabilities of titanium dioxide suspensions dispersed with acid sodium pyrophosphate or sodium pyrophosphate undergo the effects of pH, and the stabilities tend to increase with increase in pH value. In fig. 6, Tl/2 are plotted versus log concentration, C, for PA and sodium pyrophosphate at controlled pH of 6 and 10, respectively. With both PA and sodium pyrophosphate, the values of T1/2 tend to increase as pH in the systems become high values, whereas the general forms of TI/2 VS. log concentration

732
10oo l_ ' ~ ,

Colloid and Polymer Science, VoL 254 . No. 8 (1976)


~ , t '-I

0.0001

0.001

0.01

similar behaviors, although little differences between them can be found with the types of pigments. In fig. 7, the T,I2, the zeta potential, the amount of adsorption, and pH in zinc oxide suspensions are plotted as a function of the concentrations of SDS, PA and NSF, respectively. With every surfactants of SDS, PA and NSF, pH values of the zinc oxide suspensions are larger than those of the titanium dioxide
1000~
~~ 100 ~-

log concentration ofsurfactant,C(MIL)

Fig. 6. Sedimenting time Tll2 vs. log concentration curves for PA (@) and Na4P207 ( ~ ) at controlled pH of 6 ( ) and 10 ( . . . . . ). The concentration of PA refers to the concentrations of monomer unit

curves are not varied with pH, as seen in fig. 6. These may have occurred owing to the dependence of surface charge on pH and the dependence of the degree of the dissociation of materials on pH. In tab. 4, C~ and C1 obtained from log T1/2 vs. log C curves for the aqueous suspensions of inorganic pigments except titanium dioxide in solutions of various types of surfactants and inorganic phosphates are listed. This shows that the values of C~ for SDS are large in comparison with those for other surfactants, and the differences in concentration between C~ and C1 for SDS are small in comparison with those for other surfactants. The values of Cl for inorganic phosphate are small in comparison with those of the surfactants. The surfactants except SDS and inorganic phosphate listed in tab. 4 show

-f
:20

I"

=&

9 8

I 0.0001

I 0001

I 0.01

lop. concentration of surfactant, G (M/L)

Fig. 7. Sedimenting time Tl12, zeta potential, amount of adsorption, and pFI in aqueous zinc oxide suspensions vs. log concentration for SDS (O), PA (@) and NSF ((])). The concentrations of PA and NSF refer to the concentrations of monomer unit of them. A dotted and a broken line refer to the Tl12 vs. log concentration for SDS at a controlled pH of 8.9 and for p_A at a controlled pH of 10.2, respectively

Tab. 4. Ci and C1 obtained from log T,I2 vs. log C curves for zinc oxide, chrome yellow, aluminium hydroxide and day suspensions, respectively Zinc oxide C~ (M/L) PA SM OM NSF SDS TPP 0.00090 0.00040 0.00040 0.00042 0.0069 0.000082 Chrome Yellow C~ Cz (M/L) 0.00053 0.00021 0.00020 0.00042 0.0050 0.00013 (M/L) 0.011 0.0018 0.0020 0.0083 0.010 0.0014 Aluminium hydroxide C~ C1 (M/L) 0.00053 0.00021 0.00040 0.00042 0.0089 0.00013 (M/L) 0.0053 0.0016 0.0024 0.021 0.017 0.0082 Clay C~ (M/L) 0.00020 0.00021 0.00020 0.0010 0.0069 0.000082

Material

C1 (M/L) 0.010 0.0041 0.0040 0.018 0.010 0.0054

C~ (M/L) 0.032 0.0041 0.012 0.030 0.021 0.014

*) Concentration of PA, SM, OM and NSF are expressed in those of monomer units of them, respectively. TPP refers to sodium tripolyphosphate.

Moriyama, Sfabilities of aqueous inorganicpigmeni suspensions

733

suspensions. However, the relations found among stability, zeta potential, and amount of adsorption in the zinc oxide suspensions are similar as those found in the titanium dioxide suspensions. According to the D.L.V.O. theory (19), the total potential energy V between the two particles (titanium dioxide particles) is obtained by algebraic summation of the repulsive potential energy VR and the attractive potential energy VA. V = VR q- VA [2]

For a system in which aa >>1, VR given by the D . L . V . O . theory for spherical particles (19)
is: 2

V R - - ~awo 1~ (~ + e - ~ ) 2

[3]

where W is the surface potential of the particles 0 a n d / - / t h e shortest distance between the two particles. Attractive potential energy VA is given by the Hamaker equation (20), assuming that attraction between the particles is entirely based upon the London-van der Waals forces and H ~ a,

The values of V/a for aqueous titanium dioxide suspensions dispersed with 1 --1 type surfactants such as SDS and LAS, which pH of the suspensions undergo very little changes with their concentrations, have been calculated from Eq. [5] and plotted as a function of AT in figs. 8 and 9, respectively. In these figures, T~/2 vs. log C curves are shown to examine the correlation of the Via data with the stability data. The values of Via shown in figs. 8 and 9, are calculated from Eq. [5] for the concentrations of the numbers of 1, 2, 3 and 4 shown along T1/2 vs. log C curves. With SDS, Via become positive values over certain ranges o f / q for the concentrations of the numbers of 2 and 3 showing stable suspensions, as seen in fig. 8. In other words, in the systems of the numbers of 2 and 3 described above, the repulsive force works between the two pigment particles when the particles approach, and it disturbs the coagulation of the particles, leading to the stable suspensions. On the other hand, Via become negative values over all the / - / f o r the con-

1000
L

Aa 1
vA12 H [4]

100 10 1 I 0.001 t I I 0.01 0.10 log C,(M/L)

where A is the Hamaker constant. Introducing Eq. [3] and Eq. [4] into Eq. [2], we obtain.
2

via =

-T- *~* +

eWo 1 tl

e-~U)

12H

[5]
KT 5

Assuming ~00 ~ ~, and using the values of A (8.0 10-1a) calculated by Fowkes (21), the values of Via for all systems can be calculated from Eq. [5]. The value of Via depends upon the absolute value of zeta potential and the value of 1/~, i.e., thickness of the double layer. The thickness of the double layer is inversely proportional to the square of the valency of the material ion and to the concentration of the material ion, as described earlier. Accordingly, the value of Via increases with the absolute value of zeta potential and decreases with increase in the concentration of the material ion and the valency of the material ion.

A--<" H(;)

-10 Fig. 8. The relation between V/a and stability for SDS. T h e values of V/a calculated for the concentrations of the n u m b e r s of 1, 2, 3 and 4 s h o w n along r l / 2 vs. log C curves are plotted as a function of H , respectively

734

Colloid and Polymer Science, VoL 254 No. 8 (1976)


velocity data. At low concentrations the stabilities increase with the increases in concentrations of surfactants and inorganic phosphates, whereas at high concentrations they decrease remarkably with the increase in the concentrations above a certain value, which varies with the chemical composition and molecular weight of surfactants and inorganic phosphates, and also varies with the sort of inorganic pigments. The stability data at low concentrations are closely related with the zeta potential data or adsorption data, whereas the stability data at high concentrations are not related with zeta potential data or adsorption data. The stability data in the systems of 1 - 1 type surfactants such as sodium dodecyl sulfate and sodium dodecylbenzene sulfonate, in which the stabilities are little influenced by the pH changes accompanied with the increase in concentrations, can be explained by the application of the D. L. V.O. theory qualitatively. The stabilities pH-influenced, when pH values of the systems are considerably changed.

1000 2 100 3

10

o.ool I(T 5

O.Ol leg C,(M/L)

O.lO

H(1)
-5

Zusammenfassung
Es wurde der EinfluB verschiedener anionischer Tenside und anorganiscl~er Phosphate auf die Stabilit~t w~Briger anorganischer Pigmentsuspensionen untersucht. Die Stabilit~itsdaten werden aus der Sedimentationsgeschwindigkeit gewonnen. Bei niedrigen Konzentrationen nimmt die Stabilitfit mit der Konzentration der Tenside und Phosphate zu. Bei hohen Konzentrationen ist es umgekehrt. Die Stabilit~t kann nut bei niedriger Konzentration auf das Zetapotential oder Adsorptionsdaten bezogen werden. Der EinfluB der pH-Werte wird ebenfalls studiert.

-10

Fig. 9. The relation between Via and stability for LAS. The values of Via calculated for the concentrations of the numbers of 1, 2, 3 and 4 shown along TI/~ vs. log C curves are plotted as a function of H, respectively

centrations of the numbers of 1 and 4 showing unstable suspensions. In other words, in these systems, the repulsive forces do not work enough to disturb the coagulation, but the attractive force works between the two particles, leading the unstable suspensions. With LAS, similar relation between Via data and stability data can be also found, as seen in fig. 9. These results mean that the stability of aqueous titanium dioxide suspensions dispersed with 1 --1 type surfaetants such as SDS and LAS, in which pH of the suspensions undergoes very little influence with their concentrations, can be explained by applying the D. L.V.O. theory qualitatively.
Acknowledgment
The author takes this opportunity to express his thanks to Dr. T. Takeuehi and Dr. K. Nagase for their encouragements and permission to publish this paper.

References 1) Hunter, R. J. and A . E. Alexander, J. Colloid


Sci. 18, 820 (1963). 2) Mankowieh, A . M., Ind. Eng. Chem. 44, 1151 (1952). 3) Moriyama, N., K. Hattori and K. Shinoda, Nippon Kagaku Zasshi, 90, 35 (1969). 4) Morimoto, T., Nippon Kagaku Zasshi 81, 239 (1950). 5) Nakabara, Y., Nippon Kagaku Zasshi, No. 4, 695 (1972). 6) Street, N., J. Colloid Sci. 12, 1 (1957). 7) Tamamushi, B., Kolloid Z. u. Z. Polymere 150,44 (1957). 8) Tamamusbi, B. and K. Tamaki, Trans. Faraday Soc. 55, 1007 (1959). 9) Meguro, K., Kogyo Kagaku Zasshi 58, 905

(1955).
10) Takeuchi, T., F. Tokiwa, t7. Tanino and K. Haltori, Kogyo Kagaku Zasshi 74, 2239 (1971). 11) Moriyama, AT., J. Colloid lnterf. Sci. (in press). 12) Kawamura, I1. and S. Miura (to T6a G6sei Chemicals Co.), Japan, S 47--45411 13) Hattori, K. and 2". Tanino, Kogyo Kagaku Zasshi 66, 55 (1963). 14) Bossbard, H. H., R. Mory, M. Sehmid and Hah. Zollinger, Helv. Claim. Acta 42, 1653 (1959).

Summary
The effects of various types of anionic surfactants and inorganic phosphates on the stabilities of aqueous inorganic pigment suspensions have been examined. The stabilities were evaluated from sedimentation

Moriyama, Stabilities of aqueous inorganic pigment suspensions 15) Tokiwa, F. and K. Ohki, Kolloid-Z. u. Z. Polymere, 223, 38 (1968). 16) Tokiwa, F. and N. Moriyama, J. Colloid Interf. Sci. 30, 338 (1969). 17) Parks, G. A . and P. G. De Bruyn, J. Phys. Chem. 66, 967 (1962). 18) Overbeek, J. Th. G., in: Colloid Science, ed. I-I. R. Kruy~, vol. 1, pp. 207--212, p. 129 (AmsterdamHonston-NewYork-London 1952). 19) Verwey, E. jr. W. anda r. Th. G. Overbeek, in: Theory of the Stability of Lyophobic Colloids, p. 106, p. 139, (Amsterdam-New York 1948).

735

20) Iqamaker, I-I. C., Physica 4, 1058 (1937). 21) tVowkes, dv. M., in: Chemistry and Physics of Interfaces, ed. S. Ross, p. 7, (Washington, D. C. 1965).

Author's address : Noboru Moriyama Industrial Research Laboratories, Kao Soap Company, Ltd. Wakayama-shi 640--91 (Japan)

You might also like