Download as pdf or txt
Download as pdf or txt
You are on page 1of 5

Colloids and Surfaces A: Physicochem. Eng.

Aspects 541 (2018) 222–226

Contents lists available at ScienceDirect

Colloids and Surfaces A: Physicochemical and


Engineering Aspects
journal homepage: www.elsevier.com/locate/colsurfa

On the impact of surfactant type on the structure of aqueous


ferrofluids
V.I. Petrenko a,b,∗ , O.P. Artykulnyi a,b , L.A. Bulavin b,c , L. Almásy d , V.M. Garamus e ,
O.I. Ivankov a,c , N.A. Grigoryeva f , L. Vekas g , P. Kopcansky h , M.V. Avdeev a,f
a
Frank Laboratory of Neutron Physics, Joint Institute for Nuclear Research, Dubna, Russia
b
Physics Department, Kyiv Taras Shevchenko National University, Kyiv, Ukraine
c
Institute for Safety Problems of Nuclear Power Plants, Chornobyl, Ukraine
d
Wigner Research Centre for Physics, Hungarian Academy of Sciences, Budapest, Hungary
e
Helmholtz-Zentrum Geesthach, Geesthacht, Germany
f
Faculty of Physics, St. Petersburg State University, Saint Petersburg, Russia
g
Center for Fundamental and Advanced Technical Research, Romanian Academy-Timisoara Branch, Timisoara, Romania
h
Institute of Experimental Physics, Slovak Academy of Sciences, Kosice, Slovakia

h i g h l i g h t s g r a p h i c a l a b s t r a c t

• Impact of surfactant type on ferroflu-


ids structure was considered.
• Magnetic nanoparticles aggregates
of different size and type were
observed.
• Behaviour of surfactant molecules in
aqueous solutions effects MFs final
structure.

a r t i c l e i n f o a b s t r a c t

Article history: The impact of surfactant type on the structure organization of aqueous ferrofluids was considered based
Received 28 November 2016 on the data of small-angle neutron scattering (SANS). The aggregates of different sizes and types were
Received in revised form 17 February 2017 observed for magnetite magnetic nanoparticles of in heavy water stabilized by sodium oleate (SO) or
Accepted 26 March 2017
dodecylbenzene sulphonic acid (DBSA). According to the surface tension measurements it was shown the
Available online 28 March 2017
significant difference of the critical micelle concentrations for SO and DBSA solutions. The observed struc-
tural difference of the aggregates in the ferrofluids was related to the behaviour of surfactant molecules
Keywords:
Aqueous ferrofluids
Nanoparticle aggregation
Sodium oleate
Dodecylbenzene sulfonic acid
Micelle formation
Small-angle neutron scattering

∗ Correspondence to: Frank Laboratory of Neutron Physics, Joint Institute for Nuclear Research, Joliot-Curie 6, 141980 Dubna, Moscow Reg., Russia.
E-mail address: vip@nf.jinr.ru (V.I. Petrenko).

https://doi.org/10.1016/j.colsurfa.2017.03.054
0927-7757/© 2017 Elsevier B.V. All rights reserved.
V.I. Petrenko et al. / Colloids and Surfaces A: Physicochem. Eng. Aspects 541 (2018) 222–226 223

in aqueous solutions including the structure and interaction parameters of micelles (micelle aggregation
number, fractional charge, charge per micelle and surface potential, etc.) derived from the SANS analysis.
© 2017 Elsevier B.V. All rights reserved.

1. Introduction to achieve sufficient contrast for the surfactant in the SANS experi-
ment. The volume fraction of magnetite was about 0.01 and 0.013 in
Ferrofluids (FFs) or magnetic fluids are suspensions of magnetic ferrofluids with the SO and DBSA stabilization, respectively. Initial
nanoparticles with characteristic size at the level of 10 nm. The magnetite nanoparticles are quite polydisperse and its size varied
structure investigations of ferrofluids are motivated by both the from 2 up to 11 nm for both FFs [2,11,12,15]. Surfactant/magnetite
fundamental and applied interests [1–3]. Despite a wide range of mass ratio was 0.73 g/1 g in FFs with the SO stabilization. The vol-
solvents used as liquid carriers in these systems, the controllable ume fraction of the DBSA surfactant in the ferrofluid was estimated
synthesis of highly stable aqueous ferrofluids under close-to- to be about 1.7 times larger than that of magnetite.
neutral conditions is still a problem. The direct way to deal with it is In SANS experiments with pure surfactant solutions DBSA
to use steric repulsion, by coating magnetic nanoparticles (MNPs) in (CH3 (CH2 )11 C6 H4 SO3 H), and SO (CH3 (CH2 )7 CH CH(CH2 )7 COONa),
solutions by surfactant layer or layers, in order to decrease the mag- were dissolved in D2 O within a wide concentration interval of 1–15
netic dipole-dipole and van der Waals attractions between MNPs by vol.%.
increasing the mean particle-particle distance and thus preventing SANS experiments were performed at two small-angle scatter-
the aggregation in the FFs. In practice, the structure and aggrega- ing instruments including Yellow Submarine at the steady-state
tion stability of such systems are often determined by the surfactant reactor (Budapest Neutron Centre, Hungary) and YuMO at the IBR-
type and its amount in the ferrofluids. 2 pulsed reactor (JINR, Dubna, Russia) operating in time-of-flight
Thus, for FFs based on organic non-polar carriers, a single coat- regime. The isotropic differential cross-section per sample volume
ing layer of different non-saturated and saturated mono-carboxylic (hereafter referred to as scattered intensity) was obtained as a func-
acids [4,5] around magnetite MNPs can be used to provide them tion of the scattering vector module, q = (4␲/)sin(/2), where 
stability against aggregation. In this method, the stabilization effi- is the incident neutron wavelength and  is the scattering angle.
ciency and the size distribution of MNPs depends on the surfactant SANS measurements were performed at room temperature (RT).
type [6]. A correlation between the efficiency of surfactant used The fixed wavelength of 0.488 nm was used at the Yellow Sub-
for stabilization and the structure of non-polar FFs was revealed in marine diffractometer. The sample-detector distances were 1.1
the systems with surfactant excess using small-angle neutron scat- and 5.2 m that allowed to cover the range of momentum trans-
tering (SANS) [7–9]. The structure analysis was performed using fers 0.05 nm−1 ≤ q ≤ 3 nm−1 (detector size 0.64 m). The calibration
deuterated solvents, to achieve sufficient contrasts for the different on 1-mm water sample was made after the standard correc-
components of FFs. tions for background, buffer (pure D2 O) and empty cell [13]. On
For polar ferrofluids, including aqueous FFs, an excess of surfac- the YuMO small-angle spectrometer a two-detector set-up with
tants is necessary to form the second layer of surfactant molecules ring wire detectors was used. The neutron wavelength range was
on the MNP surface to provide the so-called double layer stabiliza- 0.05–0.8 nm. The measured scattering curves were corrected for the
tion. Again, in this case the surfactant type and its amount affects background scattering from buffer solutions and the absolute cali-
the ferrofluid structure and stability, as demonstrated in the pre- bration of the scattered intensity was made according to a special
vious SANS study, in which the efficiency of various surfactants to procedure using vanadium standard [14].
stabilize aqueous FFs was tested and compared [10]. The critical micelle concentration (cmc) of DBSA and SO in
The aim of the present work was to study the structure of aque- water was determined from the surface tension measurements per-
ous ferrofluids stabilized by two types of surfactants using SANS formed with the Krüss Tensiometer K20 using the ring method.
method and find out a correlation between the behavior of surfac- DBSA and SO were dissolved in bi-distilled water (Millipore) with
tants in solutions and the final structure of water-based FFs. After several concentrations within the range of 10−5 –10−3 vol. fraction.
the critical micelle concentrations were determined by surface ten- Surface tension was determined at RT by recording and averaging
sion measurements, the micelle formation of surfactants molecules ten data points at each concentration.
in aqueous solutions with and without MNPs is analyzed using the
parameters derived from the scattering curves.
3. Results and discussion

2. Experiment The experimental SANS curves for the two aqueous FFs sta-
bilized by DBSA and SO are presented in Fig. 1. Since the
The aqueous ferrofluids have been prepared by co-precipitation studied FFs are based on heavy water (scattering length density
reaction to produce nano-sized magnetite particles, and subse- SLD = 6.4·1010 cm−2 ), a scattering contribution from hydrogen-
quent addition of surfactant solutions, namely the solutions of containing surfactant (SLD ∼ 0·1010 cm−2 ) is significant due to
dodecylbenzene sulfonic acid (DBSA) and sodium oleate (SO). a high contrast. Therefore, broad bands in the curves around
Ferrofluids under the study were synthesized at the Center q ∼ 0.7 nm−1 are attributed to the scattering from the surfactants
for Fundamental and Advanced Technical Research, Romanian which form a stabilizing shell around magnetite particles, as well
Academy-Timisoara Branch, Timisoara, Romania, and at the Insti- as micelles in the solutions. It is clearly seen that the SANS curves
tute of Experimental Physics, Slovak Academy of Sciences, Kosice, for the FFs stabilized by the two types of surfactants have different
Slovakia, employing similar procedures described in [11] and [12], behavior at the small q-values. The scattering from the aqueous FF
respectively. The FFs with the SO stabilization were prepared stabilized by SO corresponds to the scattering from compact par-
directly in D2 O; in the case of the DBSA stabilization the initial FFs ticles, which is reflected in the existence of the so-called Guinier
based on H2 O was diluted (1:4) by D2 O, so the concentration of regime (the corresponding Guinier plot is given in Fig. 1) at low-
heavy water in the final fluids was 80 vol.%, which made it possible est q-values with the radius of gyration, Rg = 17 ± 1 nm. This value
224 V.I. Petrenko et al. / Colloids and Surfaces A: Physicochem. Eng. Aspects 541 (2018) 222–226

Fig. 1. Experimental SANS curves for aqueous ferrofluids stabilized by the double
layer of SO (red circle) in D2 O and DBSA (blue triangle) in the aqueous solution with
80 vol.% of D2 O. Solid lines represent the Guinier function and power-law behavior of
scattered intensity. For convenient view the curve for FF stabilized by SO is shifted
vertically by multiplying by the factor 10. (For interpretation of the references to
colour in this figure legend, the reader is referred to the web version of this article.)

Fig. 3. Experimental SANS curves from solutions of DBSA (a) and SO (b) in heavy
water with various surfactant concentrations. Arrows show the concentration
growth. Solid lines represent best fitting curves.


1 for c ≤ cmc
with g(c) = ,
0 for c > cmc
Fig. 2. Determination of the critical micelle concentration by surface tension vs
where  o is the surface tension at the surfactant concentrations
DBSA (red circles) and SO (blue triangles) volume fraction (concentration). Solid lines
show the approximations according to (1). (For interpretation of the references to higher than cmc and A = ∂/∂ln(c) (a slope in the double logarith-
colour in this figure legend, the reader is referred to the web version of this article.) mic scale) corresponds to an alteration in the surface tension as a
function of the natural logarithm of the surfactant concentration, c.
The value of minimum area per molecule at the air/water interface
is larger than that expected for separated particles, indicating the for diluted solution of monovalent surfactant can be calculated as
presence of a fraction of compact and structurally stable aggregates [16]:
in the SO stabilized FF. In the aqueous FF with the DBSA stabiliza- 2RT
tion the scattering from significantly larger aggregates is observed. S=−  ∂
, (2)
NA
This is concluded from the power-law behavior of the scattering, ∂lnC
I(q) ∼ q−2.4 , at low q-values, which indicates now a fractal-type where NA is Avogadro’s number.
organization of the aggregates with the mass fractal dimension, A large difference in the mentioned above parameters
D = 2.4. The Guinier regime is not resolved in the initial parts of was observed for the DBSA (cmc = 1.21 ± 0.01·10−2 vol.%,
the curves for these aggregates, which means that the aggregate A = −9.4 ± 0.2 mN/m and  o = 33.3 ± 0.2 mN/m, S = 90 Å) and
size, D > 120 nm (the estimate is derived from the minimum mea- SO (cmc = 3.38 ± 0.01·10−2 vol.%, A = −4.07 ± 0.1 mN/m and
sured q-value in accordance with the rule D ∼ 2␲/q), is beyond the  o = 24.7 ± 0.1 mN/m, S = 200 Å) aqueous solutions. Thus, the
instrumental limit. These observations are in full agreement with cmc-value for SO is almost three times higher than in the DBSA
the previous SANS studies of similar aqueous FFs with the double aqueous solutions. It should be mentioned that the cmc-value for
layer sterical stabilization [12,15]. DBSA is comparable with that previously obtained by the similar
The experimental values of the surface tension, (ϕ), against the method [16]. To the best of our knowledge, cmc for SO solutions
logarithm of the concentration (ϕ, vol. fraction) are plotted in Fig. 2 were obtained only by other methods but our value is comparable
for the DBSA and SO aqueous solutions. These dependences have a with that reported previously [17–19].
typical shape for micelle formation solutions described well by the The obtained experimental SANS curves from micellar solu-
expression: tions of DBSA and SO in heavy water are given in Fig. 3 a and b,
respectively. The peaks in the curves at q < 1 nm−1 correspond to
(c) = −A(ln(cmc)-ln(c))g(c) +  o , (1) the scattering structure-factor and reflect the interaction between
V.I. Petrenko et al. / Colloids and Surfaces A: Physicochem. Eng. Aspects 541 (2018) 222–226 225

Table 1
Derived from SANS data the parameters of SO micelles including micelle aggregation number, Nagg ; degree of ionization, ␣; axial ratio, ; average size, D0 ; charge, z; inverse
screening length, kd ; surface potential, 0 .

vol. fraction, % Nagg ␣ ␥ D0 , nm Z, e kd , nm−1 0 , mV

1 81(1) 0.150(1) 2.71(2) 5.27(1) 9.9(1) 0.192 189


2 96(1) 0.165(1) 2.76(2) 5.97(1) 16.4(1) 0.251 157
3 117(1) 0.155(1) 3.07(2) 6.38(1) 18.7(1) 0.298 136
4 130(1) 0.146(5) 3.02(2) 6.62(1) 19.5(3) 0.317 125
5 151(1) 0.124(5) 4.57(3) 7.96(1) 18.1(4) 0.329 100
6 181(1) 0.088(4) 4.78(2) 7.43(1) 16.3(5) 0.309 85
8 230(1) 0.070(6) 4.91(2) 8.06(1) 16.1(8) 0.322 69
9 280(1) 0.055(4) 5.62(3) 8.63(1) 16.8(7) 0.301 67
10 259(1) 0.060(3) 4.71(2) 8.40(1) 15.6(5) 0.333 60

micelles. The position of the interference maximum (q ∼ 2␲/Rint ,


where Rint is the radius of interaction) shifts to larger q-values with
increasing surfactant concentration, which points to a decrease of
the characteristic inter-micellar distances in the solution. Simi-
lar behavior for DBSA and SO solutions was previously described
[16,20] and corresponds to the charged micelles in solution.
The results of simultaneous fits of the micelle form-factor and
structure factor are shown as solid lines in Fig. 3. Spherical and
ellipsoidal form-factors were used to model the shape of micelles.
For non-spherical micelles the well-known decoupling approxima-
tion [21,22], which assumes that there is no correlation between
position and size/orientation of particles, was applied. To model
the structure factor we used the rescaled mean spherical approxi-
mation for the diluted charged colloidal dispersions developed by
Hansen and Hayter [23]. Since the studied micelles consist of the
ionic surfactant molecules, the screened Coulomb potential was
used in the model. The varied parameters included the aggrega-
tion number (Nagg ), degree of ionization (␣), axial ratio (), and
residual background were used for calculating the charge (z) of the
micelles, the inverse screening length (kd ), the surface potential
( 0 ), and the average size D0= = (3 V/4 )1/3 , where V is the volume
of a micelle. The resulting parameters for the aqueous SO solutions
are collected in Table 1. They demonstrate a quite good agreement
with our data for mixed solutions of SO with polyethylene glycol
[20] in the overlapped concentration range. The structure parame-
ters for DBSA micelles (not shown) coincide within the errors with
the previous data [16].
For the SO solutions one can see that the micelle axial ratio
grows within the interval of 2.7–5.6 with the surfactant concen-
tration, thus showing a transition towards elongated particles.
This is accompanied by an increase in the mean micelle size from Fig. 4. Aggregation number (a) and degree of ionization (b) for micelles of DBSA
D0 = 5.2 nm up to D0 = 8.6 nm. The obtained fractional charge of the (blue triangles) and SO (red circles) vs. volume fraction of surfactant. (For interpre-
micelles (about and below 0.15) is in agreement with the well- tation of the references to colour in this figure legend, the reader is referred to the
known fact that in micellar systems of ionic surfactants only a web version of this article.)
fraction of about 0.15 of counter-ions are dissociated; the rest of
them are effectively bound to the micellar surface [24]. In Fig. 4
ber of monomers of surfactants molecules (cmc-value) and MNPs
the comparison of the obtained micelle parameters of Nagg , and ␣
structure in ferrofluids, namely, the higher the cmc, the less aggre-
for DBSA and SO are presented. It can be seen that in both cases we
gated is the FF. From the structural viewpoint, the FF stabilized
have similar tendencies regarding the dependencies of the aggrega-
by DBSA is very similar to the FF modified by PEG, in which large
tion number and degree of ionization vs. surfactant concentration,
branched fractal-like aggregates were observed at addition of some
but the absolute values are slightly different. In the case of the
amount of PEG into the solutions [12]. For both ferrofluids, the
SO micelles the values of aggregation number are higher, and the
power-law behaviour at small q-values is observed together with
surfactant concentrations at which one can see the changes in the
a distinct contribution from surfactants (stabilizing shell around
power-law type behaviour of the aggregation number are shifted
magnetite particles and micelles in the solution). This is a strong
to the smaller values (Fig. 4a). These differences in the behaviour of
indication that not only the number of free surfactant molecules in
the SO and DBSA micelles can be related to the different structures
the solutions, but also the surfactant-MNPs and surfactant-solvent
of the surfactant molecules. Thus, while the overall lengths of two
interactions affect the stabilization mechanism in FFs.
surfactants are very close, the alkyl chain is branched, and the polar
head group is more complex in the DBSA molecule.
From the comparison of the behaviour of the aqueous micel- 4. Conclusion
lar solutions of DBSA and SO with the structure characteristics of
aqueous FF one can see a distinct correlation between the num- To summarize, various structure organizations of aqueous (D2 O)
ferrofluids stabilized by SO and DBSA is detected in SANS experi-
226 V.I. Petrenko et al. / Colloids and Surfaces A: Physicochem. Eng. Aspects 541 (2018) 222–226

ments. While comparatively small and compact MNPs aggregates [8] A.V. Nagornyi, V.I. Petrenko, L.A. Bulavin, et al., Structure of the
are observed in the case of the SO stabilization, large and fractal magnetite-oleic acid-decalin magnetic fluid from small-angle neutron
scattering data, Phys. Solid State 56 (1) (2014) 91–96.
type aggregates develop in water-based FF with the DBSA stabiliza- [9] V.I. Petrenko, M.V. Avdeev, V.L. Aksenov, et al., Magnetic fluids with excesses
tion. DBSA and SO micelle structures and interaction parameters of a surfactant according to the data of small-angle neutron scattering, J. Surf.
are obtained as a function of the surfactant concentration in their Invest. 3 (1) (2009) 161–164.
[10] V.I. Petrenko, V.L. Aksenov, M.V. Avdeev, et al., Analysis of the structure of
aqueous (D2 O) solutions. The determined dependences can be used aqueous ferrofluids by the small-angle neutron scattering method, Phys. Solid
in the study of complex systems with DBSA and SO where potential State 52 (5) (2010) 974–978.
presence of micelles takes a strong effect on the properties and syn- [11] D. Bica, L. Vekas, M.V. Avdeev, et al., Sterically stabilized water based
magnetic fluids: synthesis, structure and properties, J. Magn. Magn. Mater.
thesis of such systems. In particular, the micelle size is comparable
311 (2007) 17–21.
with the size of surfactant aggregates in water-based ferrofluids [12] M.V. Avdeev, A.V. Feoktystov, P. Kopcansky, et al., Structure of water-based
with the double layer stabilization. Finally, the behaviour of the ferrofluids with sodium oleate and polyethylene glycol stabilization by
small-angle neutron scattering: contrast-variation experiments, J. Appl. Cryst.
surfactants in aqueous solutions correlates with their stabilizing
43 (2010) 959–969.
properties in ferrofluids. [13] G.D. Wignall, F.S. Bates, Absolute calibration of small-angle neutron scattering
data, J. Appl. Crystallogr. 20 (1987) 28–40.
[14] A.I. Kuklin, A.Kh. Islamov, V. IGordeliy, Two-detector system for small-angle
Acknowledgements
neutron scattering instrument, Neutron News 16 (2005) 16–18.
[15] M. Balasoiu, M.V. Avdeev, V.L. Aksenov, et al., Structural organization of
The work was supported by RFBR (Russian Foundation for water-based ferrofluids with sterical stabilization as revealed by SANS, J.
Basic Research), project no. 14-22-01113-ofi m. This work was also Magn. Magn. Mater. 300 (2006) e225–e228.
[16] V.I. Petrenko, M.V. Avdeev, V.M. Garamus, et al., Micelle formation in aqueous
supported by project the Slovak Academy of Science [grant num- solutions of dodecylbenzene sulfonic acid studied by small-angle neutron
ber VEGA 2/0016/17]; Slovak Research and Development Agency scattering, Colloids Surf. A 369 (2010) 160.
[grant numbers APVV-015-0453 and APVV-14-0120]; Ministry [17] N. Mahieu, D. Canet, J.M. Cases, J.C. Boubel, Micellization of sodium oleate in
D2O as probed by proton longitudinal magnetic relaxation and self-diffusion
of Education Agency for Structural Funds of EU [grant number measurements, J. Phys. Chem. 95 (1991) 1844–1846.
26220120033]. [18] A. Hildebrand, P. Garidel, R. Neubert, A. Blume, Thermodynamics of
demicellization of mixed micelles composed of sodium oleate and bile salts,
Langmuir 20 (2004) 320–328.
References [19] N.E. Kadi, F. Martins, D. Clausse, P.C. Schulz, Critical micelle concentrations of
aqueous hexadecytrimethylammonium bromide-sodium oleate mixtures,
[1] B. V.Bashtovoi, Magnetic Berkovski, Fluids and Applications Handbook, Beggel Colloid Polym. Sci. 281 (2003) 353–362.
Inc. House, New York, 1996. [20] V.I. Petrenko, M.V. Avdeev, V.M. Garamus, L.A. Bulavin, P. Kopcansky, Impact
[2] L. Vekas, M.V. Avdeev, D. Bica, Magnetic nanofluids: synthesis and structure, of polyethylene glycol on aqueous micellar solutions of sodium oleate studied
in: D. Shi (Ed.), Nanoscience in Biomedicine, Springer-Verlag, Berlin, 2009, pp. by small-angle neutron scattering, Colloids Surf. A 480 (2015) 191–196.
650–728. [21] M. Kotlarchyk, S.H. Chen, Analysis of small angle neutron scattering spectra
[5] M.V. Avdeev, D. Bica, L. Vékás, et al., On the possibility of using short chain from polydisperse interacting colloids, J. Chem. Phys. 79 (1983) 2461–2469.
length mono-carboxylic acids for stabilization of magnetic fluids, J. Magn. [22] J. Kalus, H. Hoffmann, K. Ibel, Small-angle neutron scattering on
Magn. Mater. 311 (2007) 6. shear-induced micellar structures, Colloid Polym. Sci. 267 (1989) 818–824.
[6] M.V. Avdeev, D. Bica, L. Vekas, et al., Comparative structure analysis of [23] J.P. Hansen, J.B. Hayter, A rescaled MSA structure factor for dilute charged
non-polar organic ferrofluids stabilized by saturated mono-carboxylic acids, J. colloidal dispersions, Mol. Phys. 46 (1982) 651–656.
Colloid Interface Sci. 334 (2009) 37. [24] L.A. Bulavin, V.M. Garamus, T.V. Karmazina, et al., Measurements of structural
[7] V.I. Petrenko, M.V. Avdeev, L.A. Bulavin, et al., Effect of surfactant excess on and electrostatis parameters and surface tension of micelles of an ionic
the stability of low-polarity ferrofluids probed by small-angle neutron surfactant versus concentration, ionic strength of solution and temperature
scattering, Cryst. Rep. 61 (1) (2016) 121–125. by small-angle neutron scattering, Coll. Surf. A 131 (1998) 137–144.

You might also like