You are on page 1of 10

Surfactant Aggregation in Systems Containing

Alkanolamines and Fatty Acids


TORBJORN W,~RNHEIM ~ AND ANGELA JONSSON
Institute for Surface Chemistry, P.O. Box 5607, S-114 86 Stockholm, Sweden

Received August 24, 1989; accepted December 19, 1989

The phase diagrams of six different alkanolammonium carboxylate-water systems have been deter-
mined. Their phase behavior differs from that of ordinary soaps, since there are large differences between
the systems for varying alkyl chain lengths of the carboxylate. In addition, the chemical structure of the
alkanolammonium counterion also affects the phase equilibria. For short alkyl chain carboxylates the
liquid crystalline phases disappear (e.g., for triethanolammonium octanoate) or decrease in extension
(e.g., for monoethanolammonium octanoate), although the surfactant readily aggregates in solution. In
contrast, long alkyl chain carboxylates behave like swelling amphiphiles with a low aqueous solubility
and the immediate formation of a lamellar phase (e.g., for tri- and monoethanolammonium oleate). It
is shown that the distribution between acid and soap causes these effects. The lower surface charge of
the surfactant aggregates compared to that of ordinary soap micelles is also reflected in a lowering of the
critical micellization concentration. © 1990AcademicPress,Inc.

INTRODUCTION containing components forming mixed sur-


factant aggregates (6), i.e., alkylammonium
Fatty acids, derived from natural fats and carboxylates with two alkyl chains of similar
oils, compete with petrochemical compounds length, experimental and theoretical studies
to be raw-materials for surfactant production have explained the phase behavior in detail.
(1, 2). Today, environmental aspects have Thus it is possible to predict the phase behavior
become more and more important and this and point out the relevant interactions (7).
also implies a renaissance for soaps as such. However, for systems containing more hydro-
There is indeed a massive amount of basic re- philic amines together with the fatty acid, less
search work on soaps, at least with respect to basic understanding and information exist.
phase equilibria and association structures (3- This is so even for the common "amine soaps"
5). Perhaps the most prominent obstacle for (2), alkanolammonium carboxylates, despite
practical use of soaps, their sensitivity toward that their use as surfactants--and the use of
divalent cations, can be partly overcome in alkanolamines as additives in soap-fatty acid
technical systems by careful formulation and systems--is widespread (2, 8 ). One recent ex-
addition of complexation agents. The inter- ception is the study of the phase behavior
action between the carboxylate group and dif- and the properties of triethanolammonium
ferent ions (or other components ) in solution oleate (9-11 ).
is thus not only a fundamental research issue We have thus found it worthwhile to pursue
but also of considerable technical interest. a systematic investigation of the phase behav-
One interesting group of compounds is ior and the solution phase structure of these
mixed amine-fatty acid systems. For systems surfactants. In this paper, we give phase dia-
grams for equimolar mixtures of monoetha-
nolamine or triethanolamine and octanoic,
To whom correspondence should be addressed. dodecanoic, or oleic acid together with water.
314
0021-9797/90 $3.00
Copyright© 1990by AcademicPress,Inc.
All rightsof reproductionin any formreserved. Journal of Colloidand InterfaceScience, Vol. 138,No. 2, September1990
ALKANOLAMMONIUM CARBOXYLATE SOLUTIONS 315

The formation of liquid crystals is discussed by preparing 10-20 samples by weight in


and compared with the behavior of ordinary sealed ampoules, allowing them to equilibrate
soaps. A similar comparison is made for the at a few fixed temperatures in a thermostated
micellar aggregates formed in the solution bath, and visually determining the number
phase. The results are discussed in terms of and type of phase(s) present. A closer exam-
the differences between alkanolammonium ination was performed by mixing an addi-
ions and ordinary counterions such as Na+: tional number of samples and examining them
the amines are weak bases and they are decid- by optical microscopy, using a Reichert mi-
edly larger than alkali ions. croscope at 60× with polarizers and a fitted
hot-stage. The temperature was raised by 3 °C/
EXPERIMENTAL min. The identification of the liquid crystalline
Chemicals. The investigated surfactants phases made from microscopy (4, 5 ) was con-
were prepared from octanoic acid, Ca (Fluka, firmed by low-angle X-ray diffraction mea-
99%), dodecanoic acid, Cl2 (Aldrich, 99%), surements.
or oleic acid, C18:~ (Fluka, 99.5+%), which The determinations of ternary phase dia-
were mixed in equimolar amounts with grams were performed by titrating surfactant-
monoethanolamine (MEA) (BDH, 99.5% ), or alcohol and water-surfactant samples with
triethanolamine (TEA) (Merck, 99%). (The water and alcohol, respectively, and visually
notations MEA and TEA, respectively, are observing the phase transitions at 25 + 0.5°C.
used in the following to denote the species re- The experimental error in the boundaries of
gardless of charge.) The preparations in some the solution phases is less than 2% (w/w) and
cases were performed directly by mixing (due in the liquid crystalline phases less than
to low melting points of the products, com- 5% (w/w).
plete mixing could be attained) and in other For the binary phase diagrams the experi-
cases in solution with solvent evaporation in mental error in the solid lines is less than 2%
vacuum. No differences in melting point (or (w/w), while the experimental error in the
in other phase behavior) could be detected due dotted lines is less than 5% (w/w). The two-
to differences in methods of preparation. The phase region isotropic + anisotropic phases
compounds in the text are denoted, e.g., TEA- are, when not shown in the diagrams, judged
C~2 for triethanolammonium dodecanoate; all to be narrow.
notations are used regardless of the distribu- X - r a y measurements. Repeat distance, d,
tion of acid-soap in solution and merely to was determined with X-ray low-angle diffrac-
indicate a 1:I molar ratio of amine and fatty tion using CuK~-radiation and a camera
acid of the pure compound. Pentanol (Fluka, equipped with a Tennelec PSD-100 phase-
99%) was used as received. Water was twice sensitive detector. The temperature was 20
distilled. _+ I°C. The setup was repeatedly calibrated
There is always a risk of amidation when with crystalline sodium octanoate (d = 23.0
mixing acid and primary or secondary amines ) and the area per polar group, S, was, in
(e.g., MEA), while, of course, no such possi- the case of the tamellar phase samples, deter-
bility exists with tertiary amines (e.g., TEA). mined from the expression (4)
This reaction occurs readily at elevated tem-
peratures (around 150°C ( 1, 2)), and care was S = 2V/da, [ll
thus taken to avoid preparation or long-time where V is the amphiphile molecular volume
storage of samples at elevated temperature. (i.e., the mean value in the case of ternary
The absence of amides was verified by IR systems) and da the thickness of the amphi-
analysis of chosen samples. phile layers, obtained directly from
Phase diagrams. An overview of the phase
behavior in each binary system was obtained da = dva, [2]
Journal of Colloid and Interface Science, Vol. 138, No. 2, September 1990
316 W,~RNHEIM AND JONSSON

where va is the volume fraction of amphiphile Surface tension measurements. Surface


in the sample. All volume data were derived tension measurements were performed with
from the densities of the pure components. the de Nofiy ring method at 25 + 1°C (15).
Some complications arise when the one of the The estimated accuracy is _+0.5 m N / m .
components is distributed between the am-
phiphilic and the aqueous layers, as in the case RESULTS
of pentanol. However, the limited solubility The partial binary phase diagrams for the
of pentanol leads to maximum uncertainty investigated amphiphiles together with water
estimated at less than 2%. are shown in Figs. 1 and 2. MEA-C8 (Fig. 1a)
N M R measurements. Measurements of self- forms a solution phase L which extends up to
diffusion coefficients were performed with the 54 wt% amphiphile at 25 °C, whereafter a hex-
PGSE F T N M R method (12) on a Jeol FX- agonal liquid crystalline phase E is formed and
100 operating at 99.96 MHz for ~H, at the stable in a comparatively limited concentra-
ambient temperature 25 _+ 0.5°C. Determi- tion and temperature interval. The solution
nations of the degree ofcounterion association, phase L with MEA-C12 (Fig. lb) is stable up
/3, were performed by solubilizing trace to 30 wt% amphiphile, until a hexagonal phase
amounts of hexamethyldisiloxane (HMDS; E and, at even higher concentrations, a la-
Merck, 99%) into micellar solutions of M E A - mellar phase D form. Finally, for MEA-C18:I
C12 or TEA-C~2. The self-diffusion of H M D S (Fig. lc) the aqueous solubility of the amphi-
can be set equal to the self-diffusion of the phile is very limited and a lamellar phase D
micelle, DmiceUe( 13 ), and/3 is obtained from precipitates at higher concentrations of sur-
the measured self-diffusion of MEA or TEA, factant.
( D ) m e . . . . ed, according to TEA-C8 (Fig. 2a) forms no liquid crystals
whatsoever, and a miscibility gap occurs be-
(D}measured = /3Dmicelle+ ( 1 - 3)Dfre~, [3] tween 3 and 6 wt% at 25°C; the miscibility
where Dfr**is obtained from measurements of gap decreases somewhat with temperature. For
MEA or TEA at low concentrations in TEA-C1z (Fig. 2b) a miscibility gap occurs at
aqueous solution, corrected for the obstruction low concentrations, <0.1 wt%, extends up to
effect of the micelles (14). The error in the 7 wt% surfactant at room temperature, and
self-diffusion coefficients is less than 5% at an increases with increasing temperature. The
80% confidence interval, leading to maximum miscibility gap and other multiphase regions
errors in/3 of 10%. are denoted "undetermined" (Undet) in the

O0 100 , 100,

a b .... . . .C . . . . . . . . . L .................

80 80, 80,

?
E 6o L ~ so, L E D 60 i

Under D
40 40 , 40,

20 • • • 2O • • • 2O ° • •
2O 4O 6O 8O 100 0 2O 40 6O 80 tO0 20 4O 6O 8O
% surfactant % surfactant % surfactant

FIG. 1. Partial phase diagrams for the systems (a) monoethanolammonium octanoate-water, (b) mono-
ethanolammonium dodecanoate-water, and, (c) monoethanolammonium oleate-water. The hatched lines
denote less accurately determined phase boundaries, as stated under Experimental. Existence regions for the
solution phase L, hexagonal phase E, lamellar phase D, and solid surfactant (s) are shown. W denotes
aqueous solution dilute in surfactant. The regions of phases E and D in Figs. la and lb include the two-
phase regions E + L and D + L, respectively. The notation "Undet" is explained in the text.
Journal of Colloid and Interface Science, Vol. 138, No. 2, September 1990
ALKANOLAMMONIUM CARBOXYLATE SOLUTIONS 317

100
100
/ 1oo
a c
80 8O L

~ 6o

40
Under

/"
40
niL .J.."?sJ
y~
"d
~ 60

40 Under D

• ,...4sJ 20 ¢" 2o
20 D B U m

0 20 40 60 80 100 20 40 6O 80 100 20 40 60 80
% surfactant % surfactant % surfactant

FIG. 2. Partial phase diagrams for the systems(a) triethanolammoniumoctanoate-water(16), (b) trieth-
anolammonium dodecanoate-water( 16), and, (c) triethanolammonium oleate-water.Notations as in Fig.
1. The regions of phases E and D in Fig. 2b include the two-phase regionsE + L and D + L, respectively.

phase diagrams (Figs. lc and 2a-2c), since of 0.005 with MEA and 0.02 with TEA. These
the composition of the region varies. The values are uncertain since it is generally ac-
composition of this region in the system T E A - cepted that the effective pKa for surfactant ag-
C~2-water (Fig. 2b) has been studied in some gregates or monolayers differ from that in bulk
detail. At low concentrations of surfactant, one solution. A commonly accepted value for fatty
of the separated phases is crystalline at room acid monolayers is 5.7 (18), and for solubi-
temperature; the phase consists mainly of do- lized fatty acids in anionic surfactant aggre-
decanoic acid (m.p., 44°C). At higher con- gates, values ranging from 5.2 to 6.5 have been
centrations of surfactant, the separated phase suggested (19). Thus, our estimation is, from
is a dilute lamellar dispersion. When the tem- this point of view, a minimum value of the
perature is increased to above 40°C, the un- fraction of uncharged component.
determined region becomes an equilibrium The crucial dependence of pH on the phase
between two liquid phases. TEA-CI8:~ (Fig. behavior is verified by adding base. For ex-
2c) behaves in a manner similar to that of the ample, the system 1 M NaOH ( a q ) - T E A - C 8
corresponding MEA-salt: the aqueous solu- is a single isotropic solution phase over the
bility is low and a lamellar phase is stable at entire composition region. Adding excess base
higher concentrations ofsurfactant. Separation to the TEA-Cl2 system prevents the precipi-
of some samples in the undetermined region tation of excess acid, but the formation of a
for MEA-C~s:I and TEA-C18:1 shows the pres- lamellar dispersion is favored by adding elec-
ence of lamellar and solution phases. trolyte, as anticipated. Thus, for 1 M N a O H
We tentatively suggest--and later discuss in ( a q ) - T E A - C I 2 the miscibility gap contains
detail--that the occurrence of the miscibility liquid crystals and extends from 0.1 to 20 wt%
gaps is due to an incomplete dissociation of at room temperature but the system becomes
the acid in the systems. The pH of dilute a single solution phase region at around 45 °C.
aqueous solutions of the surfactants is above A further comparison can be made with the
7 for MEA-C8 and below 6.5 for TEA-Cs. extensively studied ternary systems water-
This should be compared with the pKa of the soap-alcohol. For water-TEA-C12-pentanol
components, which are 7.9 for TEA, 9.5 for or water-MEA-C12-pentanol, the phase dia-
MEA, and 4.8 for the acids (where the chain gram is rather similar to that for water-NaCx-
length variation of pK~ is negligible for alkyl pentanol (the main features of such phase dia-
chains longer than C6) (17). Using these grams are for ordinary soaps rather insensitive
numbers to estimate the mole fraction of un- to the value o f x (4)). An aqueous L1 and an
charged component in solutions of alka- alcohol-rich L2 solution phase occurs together
nolammonium carboxylates, we obtain values with the liquid crystalline phases E and D (Fig.
Journal of Colloid and Interface Science, Vol. 138, No. 2, September 1990
318 WARNHEIM AND J(~NSSON

C5 OH C 5 OH
b

water MEA-C12 water TEA-C12

FIG. 3. Ternary phase diagrams at 25°C for the systems (a) monoethanolammonium dodecanoate-
water-pentanol and (b) triethanolammonium dodecanoate-water-pentanol.Notations as in Fig. 1.

3). A comparison with a diagram for sodium area increases from 26 ~2 at 10 wt% water to
soap (4) shows that the diagrams are quali- 28/~x 2 at 55 wt% (Fig. 4a). The binary system
tatively similar apart from more extensive so- water-TEA-C12 behaves differently: the area
lution phases and, of course, the miscibility is constant, 41 ~ 2 between 10 and 25 wt%
gap in the water-rich region for TEA-C~2. water. The ternary system with pentanol:
From low-angle X-ray diffraction measure- TEA-C12 at a ratio 30:70 (by weight) behaves
ments, the aggregate dimensions in the la- in a manner similar to that of the MEA-C12
mellar phases have been obtained. The area system, and the area increases from 26 A 2 at
per polar group was determined for samples 25 wt% water to 29 ~2 at 55 wt% (Fig. 4b).
in the binary systems water-MEA-C~2 or - The formation of micelles and the counter-
TEA-C12 and for those in the ternary systems ion association to micelles have been briefly
with pentanol (Fig. 4). In the system water- studied. The critical micellization concentra-
MEA-C~z, the area increases from 31 fk 2 at tion, cmc, has, if possible, been determined
10 wt% water to 39 ~2 at 30 wt%. For pen- by surface tension measurements (Table I and
tanol:MEA-C12 at a ratio 30:70 by weight, the Fig. 5 ). Compared to ordinary soaps with cor-

(~2) (,~2)
a MEA b TEA
40 40

30 J 30 J 8

20 20

20 40 60 20 40 60
weight % water weight % water

FIG. 4. Areas per polar group at increasing water content for the systems (a) ©, monoethanolammonium
dodecanoate-water; e , monoethanolammonium dodecanoate-water-pentanol at constant weight ratio do-
decanoate:pentanol of 70:30 and (b) (3, triethanolammonium dodecanoate-water; e , triethanolammonium
dodecanoate-water-pentanol at constant weight ratio dodecanoate:pentanol 70:30.
Journal of Colloid and Interface Science, VoL 138,No. 2, September1990
ALKANOLAMMONIUM CARBOXYLATE SOLUTIONS 319

TABLE I TABLE II

Critical Micellization Concentrations for Different Micelle Self-Diffusion Coefficient Dmicand Counterion
Surfaetants at 25 °C Association ~ for MEA-CI2 and TEA-C~2 in Aqueous So-
lutions at 25 °C
Surfactant cmc (mol/dm 3)
Concentration Dmlc
MEA-C8 0.08 Compound (wt%) 13 (I0 n~m2 s-')

MEA-C,2 0.008
TEA-C8 0.03 MEA-C,2 10 0.65 +_ 0.05 2.6
NaC8 0.35 a 15 0.63 -+ 0.04 1.7
NaC12 0.024 ~ 20 0.62 -+ 0.04 0.9
23 0.65 +- 0.04 0.9
Value taken from Ref. (16). 25 0.63 -+ 0.04 1.0
TEA-C~2 10 0.50 -+ 0.05 2.6
15 0.54 _+ 0.05 1.9
20 0.52 + 0.03 1.3
responding alkyl chain lengths, the cmc's for 25 0.55 -+ 0.04 0.8
MEA-C8 and MEA-CI2 are lower by a factor
of approximately 3-4 (20). There is a weak
minimum in the surface tension around the
cmc for both systems (Fig. 5a), which could self-diffusion measurements (Table II). /3 is
be due either to impurities or to the small 0.60-0.65 in MEA-C12 between 10 and 25
amount of undissociated acid present. For the wt% surfactant and 0.50-0.55 in TEA-el2 in
systems with TEA-Cx, the miscibility gaps the same concentration interval. The values
prohibit measurements; however, we can in- of the micelle self-diffusion coefficient, Din,
terpolate the surface tension above and below determined fi'om the self-diffusion of the hy-
the miscibility gap with TEA-C8 (Fig. 5b), drophobic solubilizate, make it possible to es-
which suggests that aggregation occurs and timate the micelle radii from the Stokes-Ein-
that the deduced cmc is lower than that with stein equation (Table II). However, for ionic
MEA-C8 (Table I). surfactants in concentrated solutions this gives
The degree of counterion association fi in erroneous values due to the neglect of inter-
the micellar solutions of MEA-C12 and TEA- micellar interactions. Using the Dm for an or-
C12 (at concentrations above the miscibility der of magnitude estimation results in low mi-
gap, of course) have been determined from cellar radii, 10-20 A, which at least suggest

(mN/n (raN~m)
a b
60 60

50 50

4o 40

........ Nb
3o 30

20 20

i | | i,
4 3 2 1 4 3 2 1
- log c [M) -log C [M}

FIG. 5. Surface tension of solutions with varying concentration of(a) l , monoethanolammonium octanoate;
O, monoethanolammonium dodecanoate and (b) O, triethanolammonium octanoate. Miscibility gap is
indicated by dotted lines.

Journal of Colloid and Interface Science, Vol. 138, No. 2, September 1990
320 W,4,RNHEIM AND JONSSON

that no extensive micellar growth occurs; this complete association. In addition, from relax-
is supported by the concentration-independent ation measurements with an added relaxation
line width of the surfactant in the N M R agent it was concluded that the longer alkyl
spectra. chain counterions were definitely solubilized
into the micelle (23).
DISCUSSION Thus, considering the absolute values of/3
for MEA and TEA we must conclude that
Under Results we have already noted a there are no significant signs of extensive
number of differences--and similarities--be- binding due to hydrophobic or polar interac-
tween ordinary soaps and the compounds in- tions. A further argument with respect to the
vestigated here. The most drastic effect when hydrophobic interactions is that both MEA
changing counterion from, e.g., sodium to al- and TEA are highly immiscible with hydro-
kanolammonium, is that the surfactant alkyl carbon and have no hydrophobic parts, which
chain length becomes predominant in deter- facilitates penetration or solubilization into
mining the phase behavior. The phase behav- surfactant aggregates.
ior (Figs. 1 and 2) is not that of a typical ionic Let us instead consider whether the distri-
surfactant but is more reminiscent of a series bution between charged and uncharged com-
of nonionic surfactants with constant polar ponents is responsible for the effects on the
group and varying alkyl chain length. cmc's. The effects can be estimated in different
A number of factors to explain these anom- models; we use the simplest possible ideal so-
alies are considered. Not only should the in- lution theory (25-27), which has been used
teraction between the polar group of the am- to predict the cmc's of mixed surfactant sys-
phiphile and the counterion differ compared tems of ionic and nonionic surfactants with
to that in normal ions, but, in addition, hy- similar alkyl chain lengths. With this simple
drophobic interactions are often invoked in approach, it is possible to check whether the
the case of organic counterions. Both factors last hypothesis makes sense, by calculating the
should lead to a lowering of the cmc and a cmc of a hypothetical micelle with varying
significant increase in counterion association fractions of charged surfactant or vice versa.
compared to that in normal surfactant sys- We set the cmc for a completely charged
tems. In addition there is also the more trivial octanoate micelle to 0.4 M (20) and that for
fact that the acid-base equilibrium could lead a completely uncharged--a hypothetical oc-
to a neutralization of the aggregate surface. tanoic acid micelle--to 0.004 M (which equals
Turning to the hydrophobic interaction, it the solubility limit of the alcohol or the acid,
is well-known that the cmc's of a series oftet- approximately, or the cmc for an octyl ethox-
raalkylammonium dodecyl sulfates decrease ylate with a low number of ethoxy groups
with increasing size--and hydrophobicity-- (20)). Using the ideal solution expression (25-
of the alkylammonium ion (21). While we 27)
note a nonnegligible decrease in cmc with the
alkanolammonium soaps compared to that in 1/cmCmix = ( 1 -- p)/cmcionic
the corresponding alkali carboxylates (Table -[- p / CmCnonionic , [4 ]
I), the measured counterion association fl is
more similar; for sodium octanoate, fl is where p is the mole fraction of uncharged sur-
around 0.6 (22). The variation of/~ with dif- factant, we can use the cmc's of the real system
ferent organic counterions has been exten- (Table I) to calculate p within the model.
sively studied by Jansson and Stilbs (23, 24). From these values, the fraction of uncharged
For decylammonium micelles with carboxyl- component in the micelle becomes 0.04 for
ate counterions, fl can vary from around 0.50, MEA-C8 and 0.06 for TEA-C8. We can com-
indicating small interactions, to 1.0, i.e., a pare this with our previous estimation based
Journal of Colloid and Interface Science, Vol. 138,No. 2, September1990
ALKANOLAMMONIUM CARBOXYLATE SOLUTIONS 321

on the pKa's, i.e., 0.005 for MEA-C8 and 0.02 factant: the extension of the solution phases
for TEA-Cs, in dilute solutions. The discrep- (L1 and L2) decreases at the expense of the
ancy can be explained by two factors. First, multiphase region and the one-phase region
the ideal solution theory does not normally with liquid crystals (4). This explains most of
suffice to describe the data more accurately; the features of the phase diagrams in Figs. 1
instead, a regular solution theory with an and 2. In addition, the effects when comparing
added interaction parameter is used (28, 29). MEA-Cx or TEA-Cx for a fixed x are consis-
For mixtures of anionic and nonionic surfac- tent with this: the MEA carboxylate has a
tants, the interaction normally favors micel- lower fraction of uncharged component and
lization, which leads to a further lowering of behaves slightly more like ordinary soaps (the
the cmc at constant composition (28, 29). extensions of liquid crystalline phases in the
This is another way of expressing that the cmc C8 and C12 systems with respect to concentra-
is also affected by specific effects of polar tion and temperature are larger).
groups and counterions: hydration forces, hy- Only one anomaly remains: the solution
drogen bonding, and so on. Second, using a phases with both MEA-C8 and TEA-C8 are
slightly higher effective pKa of the acid as sug- anomalously extensive in the high concentra-
gested previously ( 18, 19) will make the num- tion region, even considering the undissociated
bers more self-consistent. acid. In addition, water-poor D phases nor-
We can also note that perfluorinated car- mally have high melting points even at rather
boxylate surfactants with different counterions high ratios of incorporated uncharged com-
show no corresponding drop in critical mi- ponents (4, 30), in contrast to the investigated
cellization concentration with an alkanolam- systems (Figs. 1 and 2). An increase in the
monium counterion instead of alkali ions extension of the solution phases is also noted
(29). This is because the fluorocarboxylic acids in the ternary systems water-MEA-C12 or
are stronger acids and the dissociation is com- TEA-C~2-pentanol, particularly for TEA-C12
plete even at lower pH. (Fig. 3), compared with, e.g., water-NaCx-
The phase behavior can be further illumi- pentanol (4). It has previously been shown
nated by a comparison with water-soap-al- that alkanolammonium perfluorocarboxylate
cohol or water-acid-alcohol systems, where surfactants have very low Krafft points com-
ample experimental evidence has been col- pared to the corresponding alkali compounds
lected. Phase separation may readily occur (4, (29). From Figs. 1 and 2 it is clear that this is
30) with increasing concentration of surfactant the case for hydrogenated carboxylates as well.
at a constant fraction of uncharged compo- A plausible suggestion is to ascribe the decrease
nent. At low concentrations of surfactant, hy- in melting point of the crystalline and liquid
drated and uncharged components precipi- crystalline phases compared to that in ordinary
tates; with increasing concentration of surfac- soaps to the disordering effect of the bulky
tant, there is an equilibrium between a solution counterion.
phase and a lamellar phase (4, 30). With in- Let us finally consider the aggregate dimen-
creasing temperature, the water-rich lamellar sions of the lamellar phase in the water-poor
phase melts and there is a liquid-liquid equi- region. The areas per polar group reflect a
librium or a single liquid solution phase, in packing behavior very similar to that of or-
perfect agreement with the phase diagram for dinary soaps with some differences for the bi-
TEA-Cj2 (Fig. 2b). nary water-TEA-C~2 system. For water-soap
The phase behavior can be rationalized on lamellar phases, there is normally an increase
similar grounds also at higher concentrations in area per polar group with increasing water
of surfactant. For a ternary system of water, content (4, 5 ), as is found for the water-MEA-
acid, and soap, the phase diagram is affected C~2 system (Fig. 4a). Aggregate dimensions in
by increasing the alkyl chain length of the sur- the lamellar phase can be explained with the
Journal of Colloid and Interface Science, Vol. 138, No. 2, September 1990
322 WARNHEIM AND JONSSON

J r n s s o n - W e n n e r s t r r m model (30). The elec- SUMMARY


trostatic free energy, containing the enthalpic
To conclude, alkanolammonium as coun-
(surface charge density) and entropic (mixing
terion in carboxylate systems causes nonneg-
of ions in solution) terms, balances the surface
ligible differences in phase behavior and ag-
free energy ~/A. In the binary system at low
gregation compared to ordinary soaps. The
water content, an area increase implies that
behavior in the dilute region, the lowering of
the electrostatic free energy will increase with
the c m c - - i f such exists--and the occurrence
increasing water content, which is mainly due of miscibility gaps can be ascribed to the acid-
to the decrease in entropy of mixing of the
base properties of the system; i.e., an incom-
counterions in the water lamellae at increasing
plete ionization of the acid-amine. The trends
lamellae separation. There is no increase in
of the phase diagrams with alkyl chain length
area with TEA-C12; in addition, the area is are also consistent with the fact that we have
high, 41 ~2, throughout the D-phase, which
a partitioning between acid and carboxylate
shows that the slightly lower surface charge
in the system. However, some of the noted
compared to that of MEA-C12 (or alkali car-
effects should also be ascribed to the size and
boxylates) does not cause the differences. We the character of the counterion. In particular
ascribe this to the steric interaction within the
the triethanolammonium ion has a destabiliz-
counterion layer, i.e., the size of the associated
ing effect on ordered phases due to its bulky
counterion. The radius of TEA, assuming shape, an effect useful for practical formula-
spherical shape, is around 4 A, giving a cross- tions of solution phases.
sectional area of 50 ~2, which, at these areas
per polar group and even at moderate values ACKNOWLEDGMENTS
of counterion association, leaves very little This work was financially supported by Stiftelsen Karl-
space for the TEA molecules. Adding pentanol shamns Forskning (Research Foundation of Karlshamn
ab). Many thanks to Jari Alander. Karlshamn, for dis-
to the bilayer decreases the surface charge and
cussions, to Marie Sj6berg for performing the self-diffusion
counterion association, which, together with measurements, and to Lotti Eriksson for patient technical
the additional head group spacing due to the assistance.
alcohols, leads to an area decrease as expected.
REFERENCES
The very size of TEA should thus help to
increase the curvature of the aggregates. It has 1. Falbe, J. (Ed.), "Surfactants in Consumer Products,"
been demonstrated that this factor is most of- Springer-Verlag, Berlin, 1988.
2. Johnson, R. W., and Fritz, E. (Eds.), "Fatty Acids in
ten of lesser importance and the decrease in
Industry," Dekker, New York, 1989.
curvature due to lower surface charge wins out. 3. Void, R. D., Reivere, R., and McBain, J. W., ,LAmer.
We have, though, for alkanolammonium er- Chem. Soc. 63, 1293 ( 1941 ) and references therein.
ucates, C22:1, previously noted a difference be- 4. Ekwall, P.,Adv. Liq. Cryst. 1, 1 (1975).
5. McBain, J. W., and Sierichs, W. C., J. Amer. Oil
tween MEA and TEA which could be a size
Chem. Soc. 25, 221 (1948).
effect and definitely cannot be explained by 6. Friberg, S., and Srderlfind, G., Kolloid Z. Z. Polym.
their differences in ionization (16). The M E A - 243, 56 (1971).
C22:1 and TEA-C22:l phase diagrams with wa- 7. Jokela, P., Ph.D. thesis, University of Lund, 1986.
8. "Kirk Othmer Encyclopedia of Chemical Technol-
ter are very similar to the corresponding dia-
ogy," 3rd ed. Wiley, New York, 1978.
grams for oleate. Adding MEA and TEA, re- 9. Friberg, S., Liang, P., Lockwood, F. E., and Tadros,
spectively, in excess to samples in the lamellar M., J. Phys. Chem. 88, 1045 (1984).
phase will not, at low concentrations ( ~ 10 10. Friberg, S., and Osborne D. W., Z Amer. Oil Chem.
Soc. 63, 123 (1986).
wt% excess), lead to a phase transition for
11. Friberg, S. E., Wohn, C. S., Greene, B., and van Gilder,
MEA, but the TEA system will, at excess con- R., J. Colloid Interface Sci. 101, 593 (1984).
centration, form a hexagonal phase, i.e., with 12. Stilbs, P., Prog. NMR Spectr. 19, 1 (1987).
higher curvature (16). 13. Stilbs, P., J. Colloid Interface Sci. 87, 385 (1982).

Journal of Colloid and Interface Science, Vol. 138, No. 2, September 1990
ALKANOLAMMONIUM CARBOXYLATE SOLUTIONS 323

14. Jrnsson, B., Wennerstrrm, H., Nilsson, P. G., and 22. Vikingstad, E., Skauge, A., and Hoiland, H., J. Colloid
Linse, P., Colloid Polym. Sci. 264, 77 (1986). Interface Sci. 66, 240 (1978).
15. Padday, J. F., in "Surface and Colloid Science" (E. 23. Jansson, M., and Stilbs, P., J. Phys. Chem. 91, 113
Matijevi6, Ed.), Vol. 1, p. 101. Wiley-Interscience, (1987).
New York, 1969. 24. Jansson, M., and Stilbs, P., J. Phys. Chem. 89, 4969
16. Jrnsson, A., Bokstr6m, J., Malmvik, A. C., and (1985).
W~irnheim, T., J. Amer. Oil Chem. Soc., in press. 25. Schick, M. J., and Manning, D. J., J. Amer. Chem.
17. Perrin, D., (Ed.), "Stability Constants of Metal Ion Soc. 43, 133 (1966).
Complexes," IUPAC Chem. Data Series, part B, 26. Lange, H., and Beck, K. H., Kolloid Z. Z. 251, 424
Suppl. 2. Pergamon, London, 1974. (t973).
18. Betts, J. J., and Pethica, B. A., Trans. Faraday Soc. 27. Clint, J., J. Chem. Soc. Faraday Trans. I 71, 1327
52, 1581 (1956). (1975).
19. E1 Seoud, O. A., Adv. Colloid Interface Sci. 30, 1 28. Rubingh, D. N., in "Solution Chemistry of Surfac-
(1989). tants" (K. L. Mittal, Ed.), Vol. 1, p. 337. Plenum,
20. Mukerjee, P., and Mysels, K. J., "Critical Micelle New York, 1979.
Concentrations," Nat. Stand. Ref. Data Set. Natl. 29. Kunieda, H., and Shinoda, K., J. Phys. Chem. 80,
Bureau of Standards, Washington, DC, 1971. 2468 (1976).
21. Mukerjee, P., Mysels, K. J., and Kapauan, P., J. Phys. 30. Jrnsson, B., and Wennerstrrm, H., J. Phys. Chem.
Chem. 71, 4166 (1967). 91, 338 (1987).

Journalof ColloidandlnterfoceScience,Vol.138,No. 2, September!990

You might also like