Download as pdf or txt
Download as pdf or txt
You are on page 1of 11

© 2003, e-Gnosis [online], Vol.1, Art.14 Phosphonate Surfactants. Schulz P. C.

n-Alkane phosphonic acids


The n-alkane phosphonic acids (H2CnP, in which Cn is the number of carbon atoms in the hydrocarbon
chain) have the general formula:

O

CnH2n+1⎯P⎯OH

OH

The polarity of n-alkane phosphonic acids is similar to that of aliphatic acids (Lelong and Miguens de de la
Fuente, 1974, Schulz and Lelong, 1976, Kabachnik, 1956). Table 1 shows the pKa values for the first and
second ionisation of phosphonic acids.

Table 1. pKa values for the first and second ionisation of phosphonic acids

Acid pK1 pK2 Reference _


H2C0P 1.41 6.70 a
H2C1P 2.38 7.74 a
H2C2P 2.43 8.05 a
H2C3P 2.49 8.18 a
H2C4P 2.59 8.19 a
H2C10P 3.976 ± 0.001* 7.985 ± 0.003* b
H2C12P 3.98* 8.42* c
H2C12P 2.80 8.40 a

*Thermodynamic values. References: a: Kabachnik, 1956, b: Schulz and Lelong, 1976, c: Lelong and
Miguens de de la Fuente, 1974

The crystal structure of n-decane and n-dodecane phosphonic acids was studied with some detail (Schulz,
1983). The unit cell is triclinic having four acid molecules, which are arranged in the usual structure in
amphiphile crystals: polar bilayers alternated with hydrocarbon bilayers. The crystallographic parameters are
a = 0.958 ± 0.036 nm, b = 0.538 ± 0.013 nm, c = 2.633 ± 0.028 nm, α = 93.97 ± 0.02º, β = 89.9 ± 3.3º, g =
92.93 ± 0.17º for H2C10P; and a = 0.940 ± 0.002 nm, b = 0.5963 ± 0.0007 nm, c = 3.088 ± 0.005 nm, α =
93.89 ± 0.02º, β = 81.38 ± 0.02º, γ = 79.78 ± 0.23º for H2C12P. As in other surfactant crystals, the
hydrocarbon chain structure is an all-trans zigzag conformation, naturally tilted 60º with respect to the 001
plane. This plane contains the polar head groups bilayer. The plane containing the chain zigzag is tilted with
respect to the 001 plane in an angle σ = 72.87 ± 0.26º. This gives a total angle of tilt (usually denoted by t)
of the hydrocarbon chain with respect to the 001 plane τ = 53.28º. The same value of t was found in the
crystals of their sodium salts. Because of this low value of τ, Klose et al. (Klose et al., 1982) proposed an
interdigitated structure in the crystals of the phosphonic acids that they studied (H2C7P, H2C8P and H2C10P).
ISSN:1665-5745 -2/ 16- www.e-gnosis.udg.mx/vol1/art10
© 2003, e-Gnosis [online], Vol.1, Art.14 Phosphonate Surfactants. Schulz P. C.

However, this structure leads to unbelievable values of the transversal section of the hydrocarbon chains.
The zigzag plane contains the b axis, and it is approximately perpendicular to the a axis. The hydrocarbon
chains are approximately parallel to the c axis, as it is the case in crystals of other surfactants.
3 -1
The molar volume of the phosphonic acid head group (-PO3H2) in the crystal is 16.48 ± 0.37 cm mol
(Schulz, 1983), which is smaller than that obtained in the crystals of their salts. This is explained by a strong
interaction via hydrogen bonds among neighbouring groups (both in the same layer and in the opposite layer
2
in the polar bilayer). The area occupied by a polar head group (S) in the 001 plane is 0.27358 ± 0.00033 nm ,
and is the same in acids and in the mono- and disodium salts. This situation suggest that the contraction of
the polar bilayer when passing from salts to acid is due to the approximation between the two opposite polar
layers because of the formation of hydrogen bonds. Thus, S indicates the true size of the PO3H2 group in
crystals.

When the anhydrous phosphonic acid crystals are heated, three thermal transitions appear. For H2C10P these
-1 -1
transitions occur at 43.9, 52.8 and 89.2 ºC, and the associated enthalpies are 97.5 ± 8.8 Jg , 21.3 ± 0.2 Jg
-1
and 59.0 ± 3.8 Jg respectively (Schulz et al., 1994). The first transition is a partial “melt” of the
hydrocarbon chains and a slight change in the structure and force in the polar hydrogen bonded network. The
external shape of crystals is maintained. The hydrogen-bonded network is strong enough to maintain the
structure rigid. At the polarised light microscope a smooth, waxy light grey texture is seen. Below the first
transition temperature, crystals are transparent, if they were obtained by crystallisation from aqueous or
organic solution. If they were obtained by cooling the melted acid, their texture is “craquelé”. In both cases,
crystals are strongly birrefringent. At the second transition the total melting of the hydrocarbon chains occur,
together with a change in the nature of the hydrogen bonds at the polar bilayer, which become dynamic. A
viscoisotropic (cubic) mesophase appears. Because of the strong association conserved by the polar groups
by hydrogen bonds, the cubic mesophase is probably an inverse one. All these data were obtained by FT-IR
(Schulz et al, 1996). At the third transition, the cubic liquid crystal melts giving rise to an isotropic liquid. In
saturated alkanephosphonic acid-water systems, the excess solid acid is not associated with water. All water
in the system is “bulk” or “free” water (Schulz et al., 1994), accordingly the common meaning of “bulk
water” in amphiphile-water systems (Schulz, 1998). However, when liquid crystals are present, there is some
amount of water molecules weakly associated with the polar head groups that freeze at –20 ºC with a
-1
freezing enthalpy of 47.2 ± 3.3 Jg (Schulz et al., 1994, Schulz and Puig, (1993). This water becomes free
when the acid crystallises, and the acid crystals are anhydrous.
When the acid-water system is heated the same transitions as in the anhydrous acid occur, but at lower
temperature. The thermal transitions occur at larger ranges and with lower associated enthalpies. This is
because water weakens the cohesion of the polar network structure.

Diverse liquid crystal structures may form as a function of the system water content. As an example, in the
aqueous H2C10P system an inverse hexagonal mesophase forms between 100 and 95 wt % of acid (Schulz et
al, 1996). The gel-liquid crystal transition occurs at about 15 ºC in the range 16- 80 wt % of H2C10P, giving
rise to a lamellar mesophase. The enthalpy associated with this transition decreased when the amount of
water is increased. The lamellar mesophase has different texture at the polarising microscope at low and at
high temperature, the transition between both lamellar phases occurring at about 50 ºC with a low enthalpy
change. The high temperature lamellar phase to isotropic liquid transition occurs at about 90 ºC (Schulz et
al, 1996). All these thermal transitions are of first order.

When the liquid crystal is cooled, a rigid, transparent gel forms. This gel slowly transforms into waxy solid
ISSN:1665-5745 -3/ 16- www.e-gnosis.udg.mx/vol1/art10
© 2003, e-Gnosis [online], Vol.1, Art.14 Phosphonate Surfactants. Schulz P. C.

if the temperature is above the first thermal transition, or into anhydrous crystals if the temperature is below
this transition. Schulz et al (Schulz et al, 1996) give the phase diagrams of the aqueous H2C10P and H2C12P
systems. The same article shows the typical textures of the liquid crystals. Phase diagrams of the aqueous
H2C7P and H2C8P have also been published (Klose et al., 1982), showing lamellar mesophases. However,
the same authors did not found liquid crystals in the aqueous H2C12P system (Klose et al., 1982) .

The aqueous H2CnP systems behave in a different manner than that of weakly polar amphiphiles such as
fatty alcohols and acids, which generally do not show liquid crystals (Tiddy et al., 1983), even though they
have similar polarity (Lelong and Miguens de de la Fuente,1974, Schulz and Lelong, 1976, Kabachnik,
1956).

Unlike the fatty acids, alkane phosphonic acids have critical micelle concentration (CMC), which shows
weak temperature dependence. They behave like non-ionic surfactants. The CMC values of several alkane
phosphonic acids are collected in Table 2. The Krafft point of H2C12P is 16 ºC (Minardi et al., 1996b).
-1 -1
The low surface tension at the CMC ( sCMC = 20 mN.m for H2C10P, 19 mN.m for H2C12P) (Klose et al.,
1982, Schulz et al, 1996, Yaroschenko et al., 1973) has been interpreted as an indication of strong
interaction between the –PO3H2 groups and water (Klose et al., 1982). However, the values of sCMC obtained
in solutions of extremely pure phosphonic acids were quite high (Minardi et al., 1996b). It is possible that
the low values of sCMC obtained in previous experiments arise from hydrophobic impurities, such as fatty
alcohols, as a residue of the synthesis process.

Table 2.Critical micelle concentration of alkane phosphonic acids and their mono and disodic salts.
-3
_ Critical micelle concentration / mol.dm _

Compound Temperature acid monosodic salt disodic salt Reference


ºC
H2C8P 20 0.0035 0.088 0.147 a
40 0.0035 0.093 0.177 a
60 0.0035 0.130 0.231 a
60 - 0.120 0.250 b
80 0.0035 0.0153 0.294 a
H2C9P 20 0.0027 - 0.083 a
40 0.0027 0.043 0.103 a
60 0.0027 0.056 0.153 a
60 - 0.0565 0.140 b
80 0.0027 0.069 0.143 a
H2C10P 20 0.00166 - - a
25 0.00166 c
40 0.00166 - - a
60 0.00166 0.0368 0.094 a
60 - 0.0291 0.0905 b
80 0.00166 0.0368 0.102 a
H2C12P 20 0.00104 - - a
25 0.00104 c
25 (5.4 ± 2.4)x10-4 d

ISSN:1665-5745 -4/ 16- www.e-gnosis.udg.mx/vol1/art10


© 2003, e-Gnosis [online], Vol.1, Art.14 Phosphonate Surfactants. Schulz P. C.

40 0.00104 - - a
60 0.00104 0.0055 0.0221 a
60 - 0.00563 0.0238 b
80 0.00104 0.0055 0.0221 a
H2C16P 60 0.00052 0.0013 0.0013 a
60 - 0.00117 0.00140 b
_ 80 0.00052 0.0013 0.0013 a

References: a: Demchenko and Yaroschenko,1973, b: Demchenko and Yaroschenko, 1972,


c: Yaroschenko et al., 1973, d: Minardi et al., 1996b

The critical micelle concentration of H2C12P was studied by several different methods (Minardi et al.,
1996b). Ion-selective electrode measurements indicate that the concentration of free H2C12P molecules and
+ -
free H and HC12P ions in equilibrium with micelles are nearly independent of the total concentration. H2C12P
micelles are almost uncharged and the micelle ionisation degree (a) was 0.022 at the CMC (common a
values for ionic micelles are between 0.3 and 0.5). When increasing the total surfactant concentration, a
-3
quickly decreases until a constant value of about 1.3x10 is reached in systems showing a lamellar liquid
crystal. This low a value was also found by Klose et al. (Klose et al., 1982) in other aqueous
alkanephosphonic acid systems. At the CMC an increase in the total ionisation appears, which is probably
-
due to the capture of HC12P ions by micelles. However, the first ionisation constant of the micellised acid is
-5
lower than that of the free acid (Ka1,free = 1.05x10 (Lelong and Miguens de de la Fuente,1974); Ka1,micellised ª
-6 -7
10 (Minardi et al., 1996b)). In lamellar liquid crystals, Ka1,LLC ≈ 1,8x10 (Minardi et al., 1996b). The
probable explanation of this strong reduction in the acid ionisation may be the large extension of the
hydrogen bonded polar head groups system in the Stern layer, both in micelles and lamellar liquid crystal
+
(Minardi et al., 1996b, Schulz et al, 1996, Klose et al., 1982). The H ions are probably stabilized among the
acidic oxygen atoms in the polar interface. Dimer or premicellar oligomer formation was detected at
concentrations below the CMC (Schulz et al., 1995).

The air-water interface of H2C12P solutions was studied with a combination of ion-selective electrodes,
evaporation kinetics followed with an electrobalance and surface tension measurements [(Schulz et al.,
1998). This acid forms a non-ideal monolayer below the CMC, showing a constant area per adsorbed
2
molecule (Amolec) of 0.9953 ± 0.0001 nm (Schulz et al., 1998, Lelong et al., 1976). This value is
characteristic of expanded liquid monolayers (Schukin et al., 1988, Heimenz, 1997, Sakai and Umemura,
1996). A constant Amolec value also indicates that the interface is saturated (Schukin et al., 1988). The
structural computations show that about the 65 % of the solution surface is water in contact with air. In
contrast with micelles, the polar head groups at the adsorbed monolayer are too far among them to form
2
intermolecular hydrogen bonds. Above the CMC a desorption occurs, with Amolec = 6.32 ± 0.16 nm that
remains constant with increasing concentration. This desorption may be due to energetic advantages of
micelles when compared with the adsorbed state. In micelles there is a very low contact between the
hydrocarbon chains and water, whereas in the air-water interface, 10 to 30 % of the hydrocarbon chain
length remains in contact with water (Cabane et al., 1985, Lu et al, 1992, 1993, 1993b, 1993c, 1993d, 1995,
1996 Clint, 1992, Eastone et al., 1996). The free energy change for the adsorption of a alkanephosphonic
-1
acid methylene group from bulk solution to the air-water interface is ∆µº-CH2-ads = -54,3 ± 3,6 Jmol (Lelong
-1
et al., 1976), whereas that for the micellisation is ∆µ º-CH2-mic = -161,3 ± 1,5 Jmol (Schulz and Lelong,1978).
This is an unusual situation in amphiphilic substances, because in common surfactant systems ⏐∆µºCH2-

ISSN:1665-5745 -5/ 16- www.e-gnosis.udg.mx/vol1/art10


© 2003, e-Gnosis [online], Vol.1, Art.14 Phosphonate Surfactants. Schulz P. C.

ads⏐>⏐∆µºCH2-mic⏐ (Lu et al, 1993b, Rosen, 1979). This is one of the atypical properties of this family of
surfactants. In spite of the large changes in Amolec, no significant changes in the evaporation velocity occur.
This confirms that the degree of coverage of the surface does not affect the evaporation kinetics, except if a
compact monolayer is attained (Ly et al., 1979, LaMer, 1962, LaMer and Healy, 1965, Caskey and Barlage,
1972, Mihara, 1966, McCoy, 1958).

Monosodic n-akane phosphonates

There are very scarce studies on monosodic n-alkane phosphonates (NaHCnP). The crystal structure of the
decane and dodecane homologues is known (Schulz, 1983). The anhydrous crystal density is 1109,7 ± 0,3
-3 -3
kg.m for NaHC10P and 1075,9 ± 0,4 kg.m for NaHC12P. The elementary cell is triclinc and contains four
surfactant molecules. The molecules are ordered in the common structure of organic bilayers intercalated
with polar bilayers. The crystallographic parameters are: a = 1.08 ± 0.08 nm, b = 0.634 ± 0.008 nm, c = 2.64
± 0.12 nm, α = 90.50 ± 0.08º, β = 94.0 ± 1.3º, γ = 100.03 ± 1.9º (NaHC10P) and a = 1.069 ± 0.084 nm, b =
0.633 ± 0.018 nm, c = 3.05 ± 0.11 nm, α = 91.172 ± 0.004º, β = 91.436 ± 0.004º, γ = 103.3 ± 4.3º
2
(NaHC12P). The other parameters are τ = 53.5 ± 0.1º; s = 67.996 ± 0.005º, S = 0.277 ± 0.001 nm . The polar
3 -1
group volume is 40.78 ± 0.37 cm mol . The monosodic alkane phosphonates crystallise with four water
molecules per surfactant molecule, which corresponds to the hydration of the sodium ion. This hydration
water is not detected by differential scanning calorimetry (DSC) (Soltero et al., 1998) and do not affect the
vapour pressure (Luckenheimer and Zembala, 1997, Schulz et al., 1998b). The phase diagram of the
NaHC10P – water system was published (Soltero et al., 1998). It forms micelles, a hexagonal and a lamellar
mesophases. The Krafft point of NaHC10P is 29.3 ± 1.1 ºC (Schulz and Lelong, 1979), and that of the
NaHC12P is 65 ºC (Minardi et al., 1997). The mass action model was employed to determine the micelle
+ -
ionisation degree and the concentration of micelles and free Na and HC12P ions as a function of
concentration (Schulz and Lelong, 1979). The thermodynamic parameters of micellisation of NaHC10P
(∆HºM, ∆SºM and ∆GºM) were also computed by the mass action model (Schulz and Lelong, 1979) and by
the phase separation model (Schulz and Lelong,1978).

The NaHCnP CMC slightly rises with temperature. Values for the homologues from 8 to 16 carbon atoms
are listed in Table 2. Because of the Krafft point value found for NaHC12P, the CMC at 60 ºC obtained for
the Russian authors (Demchenko and Yaroschenko, 1973, Demchenko and Yaroschenko, 1972) for this
surfactant and for NaHC16P were probably misinterpreted and are the solubility and not the CMC values.

The ionisation degree at the CMC varies almost linearly between 0.9 at 20 ºC and 1.2 at 80 ºC (Schulz and
Lelong,1978) (the maximum a value is 2). This ionisation degree was computed from the CMC dependence
on the number of carbon atoms in the hydrocarbon chain, following the theory of Molineux et al. (Molyneux
et al., 1965, Molyneux and Rodes, 1972). However, these values seem to be too large. Direct measurements
with ion-selective electrodes on aqueous NaHC12P solutions at 66 ºC found α≈ 0.12 at the CMC, which
-1 -1
increased to 0.34 at a total concentration of 0.04 mol.dm , and then decreased to about 0.08 at 0.3 mol.dm
(Minardi et al., 1997).

There is a large increase in hydrolysis upon micellisation, because of the low water solubility of the resulting
non-dissociated acid. This H2CnP is solubilised by micelles and then eliminated from the intermicellar
solution. This situation shifts the hydrolysis equilibrium to the formation of more acid (Minardi et al., 1997).
ISSN:1665-5745 -6/ 16- www.e-gnosis.udg.mx/vol1/art10
© 2003, e-Gnosis [online], Vol.1, Art.14 Phosphonate Surfactants. Schulz P. C.
-1
Between the cmc and a concentration of 0.036 mol.dm , the ratio of micellised dodecanephosphonic acid
molecule concentration to micellised monoacid dodecanephosphonate ion concentration is constant and its
- -5 -4
value is R = [H2C12P]mic/[HC12P ]mic = (1.03 ± 0.40)x10 . In comparison, R = 2,8x10 for sodium palmitate
-1 -5 -
micelles. At concentrations higher than 0.036 mol.dm R rises linearly until a value of 8x10 at 0.14 mol.dm
1
. The distribution constant of acid between micelles and the intermicellar solution KR = [H2C12P]mic/[
3 7 -1
H2C12P]water is 2x10 at the cmc and rises to 2x10 at the total concentration of 0.14 mol.dm , with a change in
-1
slope at 0.035 mol.dm . This latter concentration corresponds to a change in the structure from spherical to
wormlike micelles (Minardi et al., 1997).
3 -1
The partial molar volume of NaHC12P in water is 216.6 ± 0.2 cm mol at 66 ºC (Minardi et al., 1997). The
change in free energy of micellisation per methylene group in NaHCnP is ∆µº-CH2-mic = - (1,020 ± 0,009)kT
(Schulz and Lelong,1978), k being the Boltzmann constant and T the absolute temperature.

Disodic n-alkanephosphonates

Disodic n-alkane phosphonates have been more extensively studied than the preceding salts. In solid state
they may be found as anhydrous crystals or as tetra-hydrated crystals (Na2CnP.4H2O). Tetra-hydrated
crystals are colourless and transparent. When dehydrated, they maintain the external shape but become white
and opaque (Schulz, 1983). The four hydration water molecules cannot be freezed, as demonstrated by DSC
-3
analysis (Schulz, 1998). The Na2C12P.4H2O crystal density is 1131.2 ± 0.2 kg.m , whereas that of the
-3 -3
anhydrous crystal is 1153.6 ± 0.3 kg.m . The density of the Na2C10P crystal is 1195.5 ± 0.4 kg.m . All these
crystals have a triclinic unit cell with four surfactant molecules. The crystallographic parameters for
Na2C12P.4H2O are: a = 0.941 ± 0.57 nm, b = 0.554 ± 0.022 nm, c = 3.00 ± 0.11 nm, a = 90.124 ± 0.006º, β =
93.571 ± 0.002º, γ = 93.73 ± 0.37º. For Na2C12P they are: a = 0.967 ± 0.006 nm, b = 0.592 ± 0.004 nm, c =
3.044 ± 0.052 nm, α ≈ 90º, β ≈ 90º, γ = 83.2 ± 2.9º; and for Na2C10P: a = 0.912 ± 0.014 nm, b = 0.710 ±
0.007 nm, c = 2.654 ± 0.021 nm, α = 89.7674 ± 0.0005º, β = 88.508 ± 0.006º, γ = 94.91 ± 0.16º. The molar
3 -1
volume of the anhydrous –PO3Na2 group in crystals is de 43.08 ± 0.37 cm .mol , whereas that of the –
3 -1
PO3Na2.4H2O group is 113.49 ± 0.52 cm .mol . The area occupied by a polar group in the 001 plane is S =
2
0.22358 ± 0.00033 nm for both anhydrous and hydrated crystals. This means that the increase in volume of
the hydrated head group with respect to the anhydrous one must be due to an increase in the separation
between the two polar monolayers placed face to face in the 001 plane. The angle of tilt t is slightly smaller
in the hydrated crystal (51.1 ± 2.7º) than in the anhydrous one (53,99 ± 0,62º). However, this difference is
not statistically significant (Schulz, 1983).

The phase diagram of the Na2C10P -H2O system was published, showing the existence of a hexagonal
mesophase (Soltero et al., 1998). The micellisation of Na2CnP in water was extensively studied. The CMC’s
for several components of this family of surfactants are shown in Table 2.

The aggregation number (n) of Na2C10P micelles at the CMC is 20.8 ± 2.8 at 21ºC and 14.4 ± 0.1 at 40 ºC,
and the micelle ionisation degree (α) is 0.22 ± 0.03 and 0.032 ± 0.002 respectively. These values were
obtained from static light scattering measurements (Lelong and Schulz, 1978). For Na2C12P at 25 ºC, n =
22,5 and α = 0,223, obtained from the micelle structure analysis (Miguens de de la Fuente and Lelong,
1974). The micelle hydration is not large, 8.75 and 8.68 water molecules per micellised surfactant molecule
ISSN:1665-5745 -7/ 16- www.e-gnosis.udg.mx/vol1/art10
© 2003, e-Gnosis [online], Vol.1, Art.14 Phosphonate Surfactants. Schulz P. C.

at 21 and 40 ºC, respectively (Schulz, (1984). Since this hydration may be explained by the hydration of
counterions and the phosphonate group, the Na2CnP micelles do not have water pools in the hydrocarbon
core, as suggested by some authors (Tanford, 1974, Klotz, 1958, Fisher and Oakenfull, 1977), and
supporting the opposite opinion of Stigter (Stigter, 1974).
-
The hydrolysis increases upon micellisation, giving HCnP ions that are distributed between micelles and
-2
water with a distribution constant (for nC = 10) KR = 8.69x10 . The partial molar volume of both,
unmicellised and micellised surfactant molecules, was measured for Na2C10P and Na2C12P (Schulz, 1981).
3 -1
The partial molar volume of unmicellised Na2C10P is 195.61 ± 0.12 cm mol at 21 ºC and 206.85 ± 022
3 -1 3 -1
cm mol at 40 ºC, and 227.01 ± 0.43 cm mol for unmicellised Na2C12P at 21 ºC. For micellised Na2C10P is
3 -1 3 -1 3 -1
206.166 ± 0022 cm mol at 21 ºC and 213.256 ± 0.024 cm mol at 40 ºC, and 240.73 ± 0.15 cm mol for
micellised Na2C12P at 21 ºC. The thermal expansion coefficient of the unmicellised surfactant partial molar
-1 -1
volume is 0.003024 ± 0.000092 K , and that of the micellised amphiphile is 0.001810 ± 0.000011 K
(Schulz, 1981). This latest value is similar to that found in other micellised surfactants and in liquid
hydrocarbons (Shinoda and Soda, 1963). The methylene contribution to the micellised surfactant partial
3 -1 3 -1
molar volume was 17.28 ± 0.17 cm mol , and for the unmicellised one, 15.7 ± 0.55 cm mol . These values
agree with literature data obtained in other surfactant systems (Corkill et al., 1967, Tanaka et al., 1974). The
3 -1 3 -1
polar headgroup contribution is 38.0 ± 0.9 cm mol for micellised –PO3Na2 and 40,3 ± 2,7 cm mol for the
unmicellised one. When corrected for electrostriction and counterion volume, the partial molar volume of
2- 3 -1
the unmicellised –PO3 group is 43.22 ± 0.96 cm mol . This gives a radius of 0,235 ± 0,092 nm. The polar
headgroup size does not change upon micellisation (Schulz, 1981).
-3
Solutions below the CMC have dimmers and its formation begins at about 0.00376 mol.dm for aqueous
Na2C10P (Schulz, (1984). A study with ion-selective electrodes showed that Na2C10P micelles change their
-3
ionisation degree (α) with the total concentration, going from α = 1.15 at the CMC to 1.0 at 0,1 mol.dm and
21 ºC (Schulz, 1988/89). The maximum value of α is two. Since the value of α obtained from ion-selective
electrodes does not depend on the adopted model of the micellisation process neither on adjusted
parameters, these results suggest that the value of α = 0.22 obtained from light scattering data must be
considered with caution and is probably misinterpreted.
+ 2-
The plot of the concentration of different species in equilibrium (micelles, Na and C10P unmicellised ions) at
different total concentration is quite different to that of the majority of aqueous surfactant systems. The
+
increase in the concentration of free Na ions above the CMC is larger than expected on the basis of the
2-
common behaviour in other surfactants, but there is also an unusual increase in the free C10P ions
concentration. This situation is characteristic of systems in which the micelle structure (including size, shape
and a) is strongly dependent on the total concentration (Schulz, 1988/89). This conclusion agrees with the
unusual dependence of the Na2C12P micelle aggregation number (n) on the pH, detected by light scattering
studies. When the pH of aqueous solutions of this surfactant increases, first n diminishes and then increases
(Miguens de de la Fuente, 1973). Studies on the viscosity and hydrolysis of Na2CnCP aqueous solutions
(Schulz, 1981) and the dependence of the surface tension on concentration (Schulz, 1977) support the same
conclusion. This behaviour is due to a change in the structure of the Stern layer at the micelle surface. On
2- -
micellisation, some of the –PO3 groups become –PO3H by hydrolysis. In consequence, the surface charge
density diminishes and reduces the repulsion among polar headgroups. In addition, neighbouring headgroups
may interact by hydrogen bonding. When pH increases (by addition of NaOH or by an increase of the
ISSN:1665-5745 -8/ 16- www.e-gnosis.udg.mx/vol1/art10
© 2003, e-Gnosis [online], Vol.1, Art.14 Phosphonate Surfactants. Schulz P. C.
-
surfactant concentration), some of the OH ions overcome the electrostatic repulsion due to the negatively
- 2
charged micelle surface and reach the interfacial –PO3H groups transforming them into –PO3 ones. This
process increases the repulsion among headgroups and reduces the aggregation number. Once all the
2-
headgroups are in the –PO3 form, a subsequent increase in NaOH or surfactant concentration causes the
normal decrease in the micellised polar groups repulsion by reducing the Debye distance via the ionic
strength increase. Thus, the size of micelles increases. This effect influences the value of n, α and the CMC,
and all these phenomena were experimentally verified. The addition of NaCl to Na2C10P micelles produces
5 5
rod-like micelles with n between 1.4x10 and 3x10 (depending on the NaCl concentration), whereas the
4
addition of NaOH give micelles with n = 1,4x10 and a dependence on the added electrolyte concentration
much weaker than in the first case (Schulz, 1995). Both α and the CMC grow by addition of NaOH and
diminish on NaCl addition. The addition of a mixture of both electrolytes gives a complex response that
depends on the total concentration and the proportion of each component (Schulz, (1989).
This kind of behaviour in which the micelle properties depend on the nature and concentration of the co-ion
+ -
is rather unusual and there are some few examples in literature, associated with H y OH ions (Miguens de de
la Fuente, 1973, Schulz, 1989, 1995, Tokiwa and Ohki, 1978). In Na2CnP solutions without added
electrolyte, the increased hydrolysis produced by micellisation rises the pH. This effect increases a and
decreases n. Simultaneously, the micelle ionisation increases the ionic strength, which reduces de Debye
distance and gives rise to the opposite effects. At low total surfactant concentration predominates the first
effect, whereas the second prevails at higher concentration. As far as the author knows, only the work of
Tokiwa and Ohki (Tokiwa and Ohki, 1978) reports a similar phenomenon in a different system.

The micellisation thermodynamic parameters (∆Hºm, ∆Sºm and ∆Gºm) of aqueous Na2C10P were computed
both by the mass-action law (Lelong and Schulz, 1978) and phase separation (Schulz and Lelong,1978)
models of micellisation.

n-Alkane phosponates as components in catanionic mixtures

Catanionic amphiphiles are mixtures of a cationic and an anionic surfactant. They are a particular case of
mixed micelles. Catanionic systems are of interest in pharmacy (Zheng et al., 1983), analytical chemistry
(Zheng et al., 1983, Birch and Cockroft, 1981, Lin et al., 1983, Scowen and Leja, 1967), wastewater
treatment (URSS Patent,1983), chemical kinetic studies (Dutkiewicz et al., 1991), wetting and detergency
(Schwuger, 1971). They also are of significant interest to understand the interactions occurring in micelles
(Filipovic-Vincekovic and Skertic, 1988, Rodakiewicz-Novak, 1982, Corkill et al., 1967, Mukhayer and
Davies, 1975, Barry and Gray, 1975, Filipovic-Vincekovic, 1986, Goraleczyk, 1980, Barry and Gray, 1978).
These interactions include those between headgroups having different charge, among polar headgroups and
counterions, among different hydrophobic bodies and between the Stern layer and the neighbouring medium.
The existence in the Stern layer of headgroups having charges of opposite sign gives rise to a high charge
neutralisation and favours the formation of huge micelles (Malliaris et al., 1986). Many catanionic
equimolecular mixtures precipitate as insoluble salts (Hoyer et al., 1961).

Mixed surfactant systems show synergistic behaviour when compared with the properties of their individual
components. This synergism strongly increases in catanionic systems (Mehereteab and Loprest, 1988).

As a consequence, the understanding of the physicochemical mechanisms involved in the mixed micelle
formation, and modelling the mixed micelle formation process are of practical and theoretical interest
ISSN:1665-5745 -9/ 16- www.e-gnosis.udg.mx/vol1/art10
© 2003, e-Gnosis [online], Vol.1, Art.14 Phosphonate Surfactants. Schulz P. C.

(Rathman and Scamehorn, 1986).

Two catanionic systems involving alkanephosphonates have been studied: H2C12P-


dodecyltrimethylammonium hydroxide (DTAOH); and Na2C12P – dodecyltrimethylammonium bromide
(DTAB). Both studies gave rise to interesting conclusions on the interactions between the polar head groups
in mixed micelles. The systems were analysed by using the regular solution theory of mixed micelles (Lange
and Beck, 1973, Funasaki and Hada, 1979, Holland and Rubigh, 1983, Rubigh, 1979, Holland, 1984). This
theory studies the non-ideal interactions among different surfactant molecules in mixed micelles giving the
interaction parameter β (in kT units). This parameter is a measure of the non-ideality of interactions between
unlike surfactant molecules in the mixed micelle. Ideal interactions give β = 0 (e.g., in mixtures of
surfactants belonging to the same homologous series). Repulsive interactions give positive values of b (e.g.,
mixtures of hydrocarbon-based and fluorocarbon based surfactants (Mukerjee and Yang, 1976, Hoffmann
and Pössnecker, 1994)), and attractive interactions show negative b values. The larger the (negative)
absolute value of β, the larger the synergism of properties in the mixture is. Theoretically, β ought to be
independent on the temperature and the mixture composition (Sarmoria et al., 1992). In practice, it is slightly
temperature dependent (Rosen et al., 1994, Crisantion et al., 1994, Hey and MacTaggart, 1985, Rosen and
Zhao, 1983). In some systems it also depends on composition and an average value is often used (Crisantion
et al., 1994). However, large changes in the value of β with the system composition indicate changes in the
nature of the interactions between the components in the mixed micelles. The interaction parameter is
interpreted as (Rubigh, 1979):

β = NA(W11 + W22 – W12) (1)

where NA is the Avogradro’s number, W11 and W22 are the interaction energies between molecules in
monocomponent micelles and W12 the interaction energy between a molecule of component 1 and another of
component 2 in the mixed micelle. The β parameter has two principal contributions to the excess Gibbs’ free
energy of mixed micellisation. There is a contribution associated with the interaction between the
hydrophobic groups of amphiphiles 1 and 2, βcore, and an electrostatic contribution βelec, associated with the
interactions between the charged polar head groups of amphiphiles 1 and 2 in the Stern layer ( Rosen and
Zhao, 1983):

β ≈ βcore + βelec (2)

Some other non-electrostatic interactions between head groups (e.g., of steric, hydrogen bonding or dipolar
nature) are commonly included in the value of βelec (Moya and Schulz, 1999). Typically βcore = 0 in mixtures
having two hydrocarbon- based or two fluorocarbon-based components (Scamehorn, 1986). However, some
recent works suggest that mixtures of two hydrocarbon-based surfactants having very different hydrocarbon
backbone structure, βcore may not be zero, because of steric constraints (Proverbio and Schulz, 2002, El-Kadi
et al, 2002). Fluorocarbon-hydrocarbon mixtures give positive values of βcore (Mukerjee and Handa, 1981,
Handa and Mukerjee, 1981, Clapperton et al., 1994, Burkitt et al., 1988).

The regular solution theory also gives the mixed micelle composition and the activity coefficient of the
micellised components (γ).

ISSN:1665-5745 -10/ 16- www.e-gnosis.udg.mx/vol1/art10


© 2003, e-Gnosis [online], Vol.1, Art.14 Phosphonate Surfactants. Schulz P. C.

The aqueous H2C12P-DTAOH mixed system

The study of the mixed system H2C12P-DTAOH gave interesting information on the structure and
interactions in the micelle surface, that may be profited to design micelles having a particular surface
structure for specific applications such as micellar catalysis or analytical chemistry. The H2C12P-DTAOH
mixture is water-soluble at all proportions, even if the number of negative charges is equal to the number of
positive charges (i.e., two DTAOH molecules by each H2C12P molecule). This is an unusual behaviour
(Minardi el tal., 1996a). The system was chosen with hydrocarbon chains having the same length to obtain
βcore = 0 and polar head groups of almost the same size, to avoid steric effects arising from different head
groups size. The composition of the mixture is indicated by YH2C12P, the mole fraction of H2C12P in the
surfactant mixture (without considering water, i.e., YH2C12P + YDTAOH = 1). Between YH2C12P = 1 and 0.5 (i.e.,
between pure H2C12P and the H2C12P:DTAOH 1:1 ratio) the system behaviour is very similar to that of pure
-
H2C12P micelles. This is because the hydrogen-bonded structure among –PO3H2 and/or –PO3H groups in the
Stern layer is conserved (Schulz et al., 1998c). This was verified by determination of the local dielectric
constant at the micelle surface. In the mentioned composition range, the micelle polar surface has a local
dielectric constant (εlocal) of 60 just in the Stern layer, and exerts very scarce influence on the local dielectric
constant in the adjacent water, which reaches the pure water value of 80 very close to the micelle surface.
The β value is nearly zero, indicating a quasi-ideal interaction. When YH2C12P is diminished, β suddenly falls
up to –1.3kT at YH2C12P = 0.4. This indicates a strong attractive interaction between the two different polar
head groups. Then β slowly rises up to about –0.45kT. Simultaneously, the CMC diminishes in one order of
magnitude at YH2C12P = 0.4, and then slowly increases up to the pure DTAOH value. The micelle
composition is almost the same to that of the total composition of the amphiphile mixture up to YH2C12P =
0.7, but at higher proportions of DTAOH they are systematically richer in acid than in base. There is a sort
of “micelle azeotrope” between YH2C12P = 0.35 and 0.3, with micelles having the same composition as the
total mixture (Minardi el tal., 1996a). The study of the local dielectric constant showed that between YH2C12P
= 0.5 and 0.33 there is a very small effect of composition on the εlocal value at the Stern layer, but once
DTAOH is in excess, εlocal at the Stern layer falls to a constant value of about 28 because of the influence of
+
the –N(CH3)3 charged groups, and this influence is noticeable up to a considerable distance of the micelle
surface. In the external limit of the Stern layer, εlocal = 50 (Schulz et al., 1998c). The phenomena detected
between YH2C12P = 0.5 and 0.33 are interpreted as arising from a gradual destruction of the hydrogen-bonded
network which connects the phosphonic head groups at the Stern layer. Between YH2C12P = 0.33 and 0 (pure
DTAOH), the system behaves as a mixture of an ionic amphiphile (DTAOH) and a nonionic one (the ion
pair C12P.2DTA). Accordingly this interpretation, the micelle surface charge density (and consequently the
surface electric potential) is quite small. Micelles at the CMC have not attached counterions at the Stern
layer. When the total concentration is increased, counterions “condense” (in the sense accepted in the
polyelectrolytes’ theory (Treiner et al., 1989)) on the micelle surface. This occurs at concentrations varying
between one and three magnitude order higher to that corresponding to the aggregation of the amphiphile
ions. This phenomenon occurs even in the pure DTAOH – water system (Schulz et al., 1995).
The micelle formation in these catanionic mixtures is mainly driven by enthalpic changes. The excess
-1 -1
micellisation entropy ∆Sexess = -16.9 ± 1.2 JK mol indicates that mixed micelles are more ordered than
those of the pure components. Alternatively, it may be interpreted as that the molecular mobility in mixed
micelles is inferior to that in the pure surfactant ones (Minardi el tal., 1996a).

Since HC12P:DTA micelles behave as quasi-non ionic micelles having a very small ionisation degree, they
were used to verify that the determination of the solvent evaporation kinetics with an electronic
ISSN:1665-5745 -11/ 16- www.e-gnosis.udg.mx/vol1/art10
© 2003, e-Gnosis [online], Vol.1, Art.14 Phosphonate Surfactants. Schulz P. C.

electrobalance might be employed to determine solute molecular weights, as in static vapour pressure
measurements. The first method is more sensible than the latter (Schulz et al., 1998d). An interesting result
of this research was that the micelle formation of the HC12P:DTA-water system is not an aggregation of
monomers at the CMC, but a slow increase in the size of oligomers consisting in 3.6 ionic pairs below the
CMC to 7 ion pairs at this concentration. At the CMC a change in the aggregation mechanism occurs, and
the aggregates fast grow up to 55 or more ion pairs above this concentration.

At higher concentration, the H2C12P-DTAOH mixtures show liquid crystals. The kind of dominant
mesophase depends on the system composition. Acid-rich systems give lamellar mesophase, and DTAOH –
rich systems, hexagonal ones. At intermediate surfactants proportions there is a coexistence of lamellar,
hexagonal and cubic mesophases. Their relative amounts depend on which component is in larger proportion
(Minardi et al, 1998). Liquid crystals are richer in acid than the isotropic solutions with which they are in
equilibrium. The inclusion of DTAOH in the water- H2C12P system reduces the temperature at which liquid
crystals appear (in the pure water- H2C12P system, the lamellar mesophase appears at 30 ºC). When the
DTAOH proportion increases, hexagonal and cubic mesophases appear. The existence domains of these
liquid crystals increase and that of the lamellar mesophase diminishes when increasing the proportion of
DTAOH. Finally, only hexagonal and cubic liquid crystals appear. As in micelles, this behaviour reflects the
gradual destruction of the strong hydrogen-bonded network at the aggregates-water interface in the water-
H2C12P lamellar liquid crystal (Schulz et al, 1996).

Only the solid crystals of the H2C12P – DTAOH mixtures having excess of acid have been studied. This is
because DTAOH decomposes in concentrated solution (Morini, 1998, Jiménez-Amescua et al., 2002),
giving a Hoffmann transposition at room temperature (Jiménez-Amescua et al., 2002). Besides, crystals
obtained from their aqueous solution are inhomogeneous because they are richer in H2C12P than the initial
solution. Consequently, the successive layers formed in the crystal when it is growing vary in composition.
This is clearly visible when crystals are microscopically observed between crossed polaroids. The crystal
structure and composition of the H2C12P.0.5DTAOH mixture could be determined, because this proportion
was stable on desiccation. The elemental cell is triclinic having 18 molecules. Each half-cell has three
DTAOH molecules and six H2C12P ones. The crystallographic parameters are a = 1.27 nm, b = 2.02 nm, c =
2.22 nm, α = 120.44 ± 0.73º, β = 101.9 ± 1.3º, γ = 68.2 ± 6.2º, σ = 41.8º, τ = 35.3º. The other proportions
could not be studied by X-ray diffraction, but microscopic observation of crystals obtained by rapid
desiccation between slides (on the supposition that decomposition was not very rapid) showed that the habit
and the angles between edges were constant. In consequence, it may be concluded that the general triclinic
structure is maintained (Minardi et al, 1998). In the same article, a triangular phase diagram of this system is
shown.

The aqueous Na2C12P – DTAB catanionic system

The phase behaviour of the aqueous Na2C12P – DTAB catanionic system is much more classic that of the
H2C12P – DTAOH one (Schulz et al., 1999). The system is soluble in all proportions and four transitions
were detected when the concentration is raised. The first transition is related with a change in the state of the
-3
monolayer at the air-solution interphase, at about 0.001 mol.dm , the second is related to the formation of
-3
ion pairs at about 0.0065 mol.dm . The third transition is the CMC, and the fourth to the spherical to rod-
like micelles transformation. The two latter concentrations depend on the Na2C12P – DTAB proportion. The
interaction between the two amphiphilic salts in micelles is weaker than that observed in other catanionic
systems formed with salts, with b = (-1.66 ± 0.05)kT. The interaction parameter shows a weak dependence
ISSN:1665-5745 -12/ 16- www.e-gnosis.udg.mx/vol1/art10

You might also like