Sei sulla pagina 1di 9

Colloids and Surfaces A: Physicochem. Eng.

Aspects 436 (2013) 675–683

Contents lists available at ScienceDirect

Colloids and Surfaces A: Physicochemical and


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

Surfactant effects on the microstructures of Fe3 O4 nanoparticles


synthesized by microemulsion method
Ting Lu a , Junhu Wang a,b , Jie Yin a,c , Aiqin Wang a , Xiaodong Wang a,∗ , Tao Zhang a,b,∗
a
State Key Laboratory of Catalysis, Dalian Institute of Chemical Physics, Chinese Academy of Sciences, 457 Zhongshan Road, Dalian 116023, PR China
b
Mössbauer Effect Data Center, Dalian Institute of Chemical Physics, Chinese Academy of Sciences, 457 Zhongshan Road, Dalian 116023, PR China
c
Graduate University of Chinese Academy of Sciences, Beijing 100049, PR China

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

• A series of Fe3 O4 nanoparticles have


been synthesized by microemulsion
method.
• Crystal lattice defect and magnetite
stoichiometry are reported.
• Variation of surfactant structure
affects the characteristics of Fe3 O4
nanoparticles.

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

Article history: A water-in-oil microemulsion route has been devised to synthesize nanosized magnetite (Fe3 O4 ) parti-
Received 3 June 2013 cles using different kinds of surfactant as the surfactant phase, n-heptane as the oil phase, and n-hexanol
Received in revised form 1 August 2013 as the co-surfactant phase, respectively. The X-ray diffraction (XRD), transmission electron microscopy
Accepted 4 August 2013
(TEM) and 57 Fe Mössbauer technique were employed to investigate the characteristic of the nanoparti-
Available online xxx
cles. The Fe3 O4 powder derived from various microemulsions possessed an average spherical particle size
of 13–15 nm based on TEM observation. The 57 Fe Mössbauer spectroscopy results reveal that surfactant
Keywords:
structure plays an important role in regulating the microstructure of Fe3 O4 nanoparticles. The relation-
Surfactant effects
Fe3 O4 nanoparticles
ships among the crystal lattice defect (vacancy parameter), magnetite stoichiometry and the surfactant
Microstructures structure have been discussed in terms of headgroup charge, hydrophobic chain length, headgroup size.
Microemulsion The temperature effects on the particle size and defect are performed and the most remarkable factor of
surfactant structures on the lattice defect and magnetite stoichiometry against the temperature variation
is also discussed. Moreover, the influence of lattice defect and magnetite stoichiometry on the magnetic
property has also been explored.
Crown Copyright © 2013 Published by Elsevier B.V. All rights reserved.

1. Introduction application in biology, medicine and catalysis areas [2–7]. In addi-


tion, as a functional material, magnetic nanoparticles of Fe3 O4 is
As an important iron oxide [1], magnetite (Fe3 O4 ) has been also used for recording materials, magnetic fluid materials and elec-
intensively concerned and researched in recent years owing to tronic materials, etc. [8–10]. It has been realized that the properties
its special properties in chemistry and physics as well as wide of nanosized functional particles are very different from those of
bulk materials in the same composition [11]. Shape, size and size
distribution of the nanoparticles are the key factors that determine
their chemical and physical properties, so as to their functionality
∗ Corresponding author at: State Key Laboratory of Catalysis, Dalian Institute
[12–14]. As an example, nano-Fe3 O4 has shown some features not
of Chemical Physics, Chinese Academy of Sciences, 457 Zhongshan Road, Dalian
116023, PR China. Tel.: +86 411 84379015; fax: +86 411 84685940.
available for the bulk Fe3 O4 , such as superparamagnetism, small-
E-mail address: taozhang@dicp.ac.cn (T. Zhang). size effect and quantum-tunnel effect. Therefore, preparation of

0927-7757/$ – see front matter. Crown Copyright © 2013 Published by Elsevier B.V. All rights reserved.
http://dx.doi.org/10.1016/j.colsurfa.2013.08.004
676 T. Lu et al. / Colloids and Surfaces A: Physicochem. Eng. Aspects 436 (2013) 675–683

such functional nanoparticles involves a very good understanding O


of the influence of synthesis parameters on the products properties, O S O Na N Br
(a) SDS Br N
such as their size, morphologies, crystal defects, etc. O
Currently, various routes and methods, with each advantage and
(OCH2CH2)4OH
disadvantage, have been developed to synthesize magnetic Fe3 O4 (b) Brij30
nanoparticles, such as mechanical attrition method, microemulsion
method, hydrothermal method, co-precipitation method [15–18]. N
(c) DTAB Br
Among these synthesis methods, microemulsion method is a novel
method for nanoparticles preparation and has been widely used
N
to synthesize various kinds of nanomaterials in recent years, (d) DEAB Br
such as catalytic and magnetic materials [16,19–22]. Microemul-
sions are thermodynamically stable dispersions of immiscible
water phase and oil phase stabilized by the arrangement of sur- N
(g) 12-2-12
factant and co-surfactant molecules at the interface [23]. The (e) DBAB Br
microemulsion, especially water-in-oil (W/O), which consists of
water nanodroplets dispersed in the bulk oil phase and is stabilized N
in spherical reverse micelles, can be considered as “nanoreactor” (f) CTAB Br
for the synthesis of nanoparticles. When two water nanodroplets
Fig. 1. Chemical structures and corresponding abbreviations of various surfactants
collide, which are driven by the Brownian motion, they fuse and used in microemulsions: (a) sodium dodecylsulfate, (b) polyoxyethylene(4) lau-
interchange reactants [24]. Since the chemical reaction is limited to ryl ether, (c) dodecyltrimethylammonium bromide, (d) dodecyltriethylammonium
the nuclear interior in water, the microemulsion method has obvi- bromide, (e) dodecyltributylammonium bromide, (f) cetyltrimethylammonium bro-
ous advantages to manipulate the particles size and shape, also to mide, (g) dimethylene-1,2-bis(dodecyldimethylammonium bromide).
regulate narrow particle size distribution by controlling the amount
of the mixture oil/water/surfactant [25,26]. were purchased from Acros Organics Co. and used as received.
By microemulsion dynamics, we note the fact that each of water Quaternary ammonium bromides including dodecyltriethylam-
nanodomain is not static one, but in continuous movement and monium bromide (DEAB), dodecyltributylammonium bromide
collision with each other [24]. In each collision, material inter- (DBAB) were prepared by the reaction of corresponding alkyl
change can take place. This phenomenon is called intermicellar bromide and trialkylamine. Gemini surfactant dimethylene-1,2-
exchange which is strongly dependent on the elasticity of the bis(dodecyldimethylammonium bromide), abbreviated as 12-2-12,
surfactant film [27]. Classically, the microemulsion exchange char- was synthesized according to the literature method [29]. All the
acteristic time ( ex ) is in the range of 10 ␮s <  ex < 1 ms, depending synthesized surfactants were recrystallized five times from mixed
on the film flexibility which is mainly determined by the used solvents of acetone-ether or ethanol-ethyl acetate. The purity of the
surfactant [24]. On the other hand, the characteristic of crystal- surfactants was examined, and no surface tension minimum was
lite, especially for lattice defect, is mainly controlled by nucleation found in the surface tension curve.
and subsequent growth rate from the perspective of crystal forma-
tion kinetics [16]. Thus, the defects in crystallite are closely related
2.2. Preparation and purification
to the process of reactants to be exchanged, i.e. exchange char-
acteristic time ( ex ), corresponding to each surfactant used in the
The Fe3 O4 nanoparticles were prepared via the following reac-
microemulsion. Therefore, to further understand the synthesis of
tion equation:
nanoparticles by microemulsion method, there is a need to know
the influence of microemulsion dynamics on the crystallinity of 8OH− + Fe2+ + 2Fe3+ → Fe3 O4 ↓ +4H2 O (1)
nanoparticles, i.e. surfactant structures effect on the crystal char-
acteristics of nanoparticles. However, so far few studies [28] have In a typical synthesis, 2.0 g surfactant was dissolved in 20 mL n-
been focused on the related work. heptane and 12 mL n-hexanol to obtain a transparent solution by
In this work, magnetic nanoparticles of Fe3 O4 were synthesized ultrasound or stirring. The above solution was then put into a
from water-in-oil microemulsions used different kinds of surfac- three-necked flask and stirred for 20 min at 40 ◦ C under an argon
tants (Fig. 1), including conventional single-chain surfactants and atmosphere. 0.5 mol/L of freshly prepared aqueous iron(II) sul-
novel Gemini surfactant. The morphology, size, size distribution fate in 0.6 mL double distilled water was then injected rapidly.
and crystalline feature of the nanoparticles were investigated. In After 2 min, 1.0 mL aqueous ferric chloride solution (0.5 mol/L) was
particular, the effects of various surfactant structures, such as the added. The solution turned into light brown color. The emulsion
charge of polar headgroup, the hydrophobic chain length and the containing 2 mL ammonia water (25% NH3 ) as the water phase
hydrocarbon parts of polar headgroup, on the lattice defect and was also prepared using the same oil and surfactant phase. After
magnetite stoichiometry of Fe3 O4 nanoparticles were discussed in 20 min, the emulsion containing ammonia was injected dropwise
detail. Meanwhile, this work can lay a foundation in studying the into the above three-necked flask with vigorous stirring at 70 ◦ C.
relationship between the surfactant structures as well as the prop- The color of the solution quickly turned from light brown to black.
erties of synthesized nanomaterials by microemulsion method. The entire solution was heated at 70 ◦ C for 3 h. The whole procedure
was carried out under an argon atmosphere to prevent the oxida-
tion of ferrous ions. Finally, the crude product was aged for 2 h at
2. Experimental room temperature before being repeatedly washed with ethanol
and water. The crude product was then retrieved by centrifugation
2.1. Reagents and material and dried in a vacuum oven at 80 ◦ C for 8 h.

The starting chemicals used in this work were: iron(II) 2.3. Characterization
sulfate heptahydrate (FeSO4 ·7H2 O), ferric chloride hexahydrate
(FeCl3 ·6H2 O) and ammonia solution (25% NH3 ) of A.R. grade (Tian- The X-ray diffraction (XRD) patterns were recorded with a PAN-
jin Kermel Co., China). Surfactants SDS, Brij30, DTAB, and CTAB alytical X’pert Pro Super model diffractometer with Cu K␣ radiation
T. Lu et al. / Colloids and Surfaces A: Physicochem. Eng. Aspects 436 (2013) 675–683 677

( = 1.5432 Å) at a scan speed of 2 ◦ /min. The transmission elec- Table 1


Average crystal sizes (d), lattice constants (a, b, c), average particle diameters (D)
tron microscopy (TEM) observation was performed on a JEOL 2000
and standard deviations (SD) of Fe3 O4 nanoparticles calculated from XRD patterns
EX electronic microscope with an accelerating voltage of 120 kV. and TEM observation.
57 Fe Mössbauer spectra of the as-prepared materials were recorded
Surfactants d (nm)b a = b = c (nm)c D (nm)d SD (nm)
using a Topologic 500A spectrometer and a proportional counter
at room temperature. The source was 57 Co in rhodium. Each of SDS 14 0.8388 13.0 1.7
Brij30 16 0.8372 14.8 2.6
the Mössbauer spectra was fitted by two overlapping six-line sub-
DTAB 14 0.8371 14.2 2.2
spectra. One of the sub-spectra with larger magnetic splitting was DEAB 15 0.8369 14.2 2.0
assumed to have a static magnetic field, the other one with smaller DBAB 14 0.8370 14.2 2.2
magnetic splitting was assumed to have a distribution in its mag- CTAB 14 0.8375 14.2 2.0
netic field. Accordingly, 57 Fe Mössbauer spectral parameters such 12-2-12 15 0.8379 14.6 2.3
STANDARDa – 0.8375 – –
as the isomer shift (IS), the electric quadrupole splitting (QS), the
a
effective or average magnetic field (H), the full linewidth at half Standard pattern from JCPDS No. 1-88-315.
b
Calculated by Scherrer formula.
maximum (FWHM) and the relative spectral area (A) of different c
Calculated through MDIJade5.0 software.
components on the absorption patterns were calculated. The IS d
Calculated through Smile View software (Version 2.05) based on TEM micro-
values were quoted relative to ˛-Fe at room temperature. The mag- graphs.
netization of the samples was measured with a vibrating sample
magnetometer (Lakeshore730, USA) at room temperature.
Comparing with the standard pattern of Fe3 O4 power, the peaks of
our samples were clearly broadened, implying the decrease of the
3. Results and discussion
particle size [30]. Using Scherrer equation [31] for the FWHM of
these obvious diffraction traces (2 2 0), (3 1 1), (5 1 1) and (4 4 0),
3.1. XRD and TEM characterization
the average crystallite sizes are calculated and listed in Table 1,
and also the lattice constants of Fe3 O4 nanoparticles are shown. It
The XRD patterns of the powders derived from the water-
can be seen from Table 1, the particle sizes by different surfactant
in-oil microemulsions are shown in Fig. 2. The diffraction peaks
systems were close to each other of 14–16 nm, and the calculated
were characteristic of the cubic spinel-type Fe3 O4 without other
lattice constants of our samples were consistent with the standard
iron oxide phases, such as ␣-Fe2 O3 and ␥-Fe2 O3 , indicating that
reverse spinel-type Fe3 O4 .
single-phase Fe3 O4 can be obtained by our microemulsion method.
Fig. 3 shows bright field TEM micrographs for the Fe3 O4 powders
derived from microemulsions of various surfactants. The images
(311)

showed that in despite of a few irregular shapes, all samples


were nearly spherical particles going with slight distortion and
aggregation to some extent. For all the samples, the nanoparticles
synthesized in SDS microemulsion showed better dispersion than
(440)
(220)

(511)

other surfactants. The size distribution of Fe3 O4 nanoparticles were


(400)

(422)

calculated at least 400 particles by image analysis software, which


is also illustrated in Fig. 3. It can be seen that the size distribution of
SDS most particles was among the range of 10–20 nm for all the samples,
together with a small quantity of particles exhibiting a relatively
Brij30 wider size distribution larger than 20 nm. Also, the average particle
diameters and standard deviations derived from size distribution
histograms are listed in Table 1, which were consistent with the
DTAB calculated results of Scherrer equation from XRD data. The regular
Intensity

particle size and morphology of the Fe3 O4 nanoparticles derived


from various surfactant microemulsions can be attributed to the
DEAB fact that the nucleation and subsequent growth of Fe3 O4 nanopar-
ticles took place on a controlled manner in the nanodomains of
DBAB water phase, which were dispersed in a continuous oil phase with
surfactant molecules residing at the interface. It indicated that the
unique synthesis process of microemulsion method brought an
CTAB improvement in the uniformity of crystallite size and morphology.
Summary of the previous sections, nanosized Fe3 O4 particles
have been synthesized by a water-in-oil microemulsion route using
12-2-12 several kinds of surfactants, respectively. By XRD and TEM charac-
terization methods, only a few surfactant effects can be observed on
Fe3O4 the crystalline phase, size and morphology of Fe3 O4 nanoparticles
synthesized by microemulsion method using different surfactants.
-Fe2O3
3.2. 57 Fe Mössbauer characterization

20 30 40 50 60 70 80 Mössbauer technique is effective to identify the iron state in


the functional materials; it has been widely employed to study
2 (degree) fine magnetic particles and has given much unique information
Fig. 2. XRD patterns of Fe3 O4 nanoparticles synthesized in microemulsions of var-
about the electronic, structural and magnetic properties of such
ious surfactants at 70 ◦ C and also including the standard patterns of Fe3 O4 and systems [32–36]. In order to further explore the surfactant effects
␥-Fe2 O3 . on the microstructure of Fe3 O4 nanoparticles, 57 Fe Mössbauer
678 T. Lu et al. / Colloids and Surfaces A: Physicochem. Eng. Aspects 436 (2013) 675–683

Fig. 3. TEM micrographs and size histograms for Fe3 O4 nanoparticles prepared by microemulsion method at 70 ◦ C in (a) SDS, (b) Brij30, (c) DTAB, (d) DEAB, (e) DBAB, (f)
CTAB, (g) 12-2-12. The scale bar is 20 nm.

spectroscopy was employed to characterize the occupied states arises from a superposition of contributions by Fe3+
A
(component
of irons in the crystal and the deviation of magnetite from sto- A) and the second one from a mixture of Fe2+ and Fe3+ (component
A B
ichiometry. Fig. 4 shows the room temperature 57 Fe Mössbauer B), respectively.
spectra of the as-prepared Fe3 O4 nanoparticles by water-in-oil Table 2 lists the obtained 57 Fe Mössbauer spectral parameters.
microemulsion using different surfactants. Data can be seen from Table 2: (1) the IS values of the iron ions in B
In order to interpret the observed Fe3 O4 spectra, it is useful to sites (Fe2+ and Fe3+ ions in octahedral sites) were higher than those
consider the cation distribution. The structure of magnetite is an of A sites (Fe3+ ions in tetrahedral sites), which can be attributed
inverse spinel which belongs to cubic crystal system. As known, its to the decision of the valence state of iron ions; (2) QS values of
formula can be written in the general form as Fe3+ [Fe2+ Fe3+ ]O4 with all samples were close to zero, indicating the charge distribution
Fe cations without brackets on tetrahedral (A) sites and those in at the iron nuclear position in Fe3 O4 samples is very symmetric;
brackets on octahedral (B) sites of the reverse spinel lattice [35,37]. (3) the magnetic field intensities of iron ions in octahedral (B) sites
Magnetite can have a range of oxidation states dependent upon the were smaller than those of tetrahedral (A) sites; (4) nearly all the
amount of structural Fe2+ , which can be discussed quantitatively line width of the spectra was smaller than 0.582 mm/s.
as the magnetite stoichiometry (x = Fe2+ /Fe3+ ). For magnetite with Further observation showed that the main difference of samples
an ideal Fe2+ content (assuming the Fe3 O4 formula), the mineral in Mössbauer spectra was the relative area of A and B sites. Since the
phase is known as stoichiometric magnetite (x = 0.50). As magnetite peak area ratios (SB /SA ) of different samples respond to the ratios
becomes oxidized, the Fe2+ /Fe3+ ratio decreases (x < 0.50), with this of iron ions in octahedral (B) sites and tetrahedral (A) sites of the
form denoted as nonstoichiometric or partially oxidized magnetite. Fe3 O4 crystal, thus they reflect the differences in structure and com-
When the magnetite is completely oxidized (x = 0), the mineral is position. For nonstoichiometric magnetite, the structure is often
known as maghemite (␥-Fe2 O3 ). Magnetite stoichiometry has been written as Fe3−ı O4 or expressed as Tet Fe3+ [Oct Fe2+ Fe3+  ]O4
1−3ı 1+2ı ı
extensively studied; however, only little work explored the stoi- to express the composition of the sample, where ı is the corrected
chiometry of nanoparticle samples. As shown in Fig. 4, all spectra value within the limits of 0 (stoichiometric magnetite) and 1/3
can be made of two six-line patterns fitting, of which one sextet (completely oxidized, i.e. stoichiometric maghemite ␥-Fe2 O3 ), 
T. Lu et al. / Colloids and Surfaces A: Physicochem. Eng. Aspects 436 (2013) 675–683 679

Table 2
Room temperature 57 Fe Mössbauer parameters of Fe3 O4 nanoparticles synthesized in microemulsions of various surfactants at 70 ◦ C.

Surfactants Sub-spectra Oxidation state of iron ISa (mm/s) QSb (mm/s) Hc [T] Spectral area (%)d
2+ 3+
SDS Octahedral Fe /Fe 0.41 0.02 35.2 87.6
Tetrahedral Fe3+ 0.28 0.02 47.7 12.4
Brij30 Octahedral Fe2+ /Fe3+ 0.41 −0.02 41.1 82.4
Tetrahedral Fe3+ 0.28 0 48.1 17.6
DTAB Octahedral Fe2+ /Fe3+ 0.48 0 44.6 81.5
Tetrahedral Fe3+ 0.28 −0.04 47.1 18.5
DEAB Octahedral Fe2+ /Fe3+ 0.44 −0.01 38.5 78.4
Tetrahedral Fe3+ 0.27 −0.01 47.6 21.6
DBAB Octahedral Fe2+ /Fe3+ 0.50 −0.01 41.3 73.3
Tetrahedral Fe3+ 0.28 −0.02 47.7 26.7
CTAB Octahedral Fe2+ /Fe3+ 0.39 −0.02 39.3 74.7
Tetrahedral Fe3+ 0.29 0.01 47.7 25.3
12-2-12 Octahedral Fe2+ /Fe3+ 0.43 0 41.7 67.0
Tetrahedral Fe3+ 0.30 0 48.5 33.0
a
IS, isomer shift, related to ␣-Fe.
b
QS, electric quadrupole splitting.
c
H, magnetic field.
d
Uncertainty is ±5% of reported value.

Table 3
Area ratios of peak B and peak A in Mössbauer spectra (SB /SA ), vacancy num-
bers (ı) and magnetite stoichiometry (xMS ) of Fe3 O4 nanoparticles synthesized in
microemulsions of various surfactants at 70 ◦ C.

Surfactants SB /SA |ı| xMS

SDS 7.07 0.1226 0.282


Brij30 4.68 0.0912 0.333
DTAB 4.41 0.0858 0.342
DEAB 3.63 0.0675 0.374
DBAB 2.75 0.0378 0.427
CTAB 2.95 0.0459 0.412
12-2-12 2.03 0.0019 0.496

1 − 3ı
xMS = (3)
2 + 2ı
Here, SB /SA is the area ratio of peak B and peak A in Mössbauer
spectra, which stands for the ratio of iron ions in octahedral to
tetrahedral sites, because the intensity of a Mössbauer spectrum is
proportional to the number of corresponding Mössbauer nuclei in
the sample. The absolute value of ı is used in Eq. (3) to calculate the
magnetite stoichiometry. In addition, we assumed that the recoil-
free fraction of tetrahedral (A) sites is equal to that of octahedral
(B) sites.
The ı and xMS values of the samples derived form various sur-
factant microemulsions are calculated and listed in Table 3. The
knowledge of the deviation from stoichiometry is important, which
Fig. 4. Room temperature 57 Fe Mössbauer spectra for Fe3 O4 nanoparticles synthe-
can dramatically influence the particles’ physical and chemical
sized in microemulsions of various surfactants at 70 ◦ C: (a) SDS, (b) Brij30, (c) DTAB,
(d) DEAB, (e) DBAB, (f) CTAB, (g) 12-2-12. properties, particularly in some problems of oxidation and corro-
sion of iron and steel. Mössbauer spectroscopy is considered as
a useful technique to obtain the information on the nonstoichio-
metry of magnetite, especially in a very small quantity of product
are vacancies formed in the crystal structure to account for charge [38]. Seen from Table 3, the ı values were calculated to be negative
balance. In order to obtain the surfactant effects on the micro- ones for all the samples, indicating that the ferrous ions were in
structures of Fe3 O4 crystal, two equations [32,38] as followed were excess or oxygen atoms were in short comparing with stoichiom-
applied to determinate the vacancy parameter (ı) and the devia- etry magnetite [39]. As shown in Table 3, the absolute values of ı
tion of magnetite from stoichiometry (xMS ), namely, differences in varied from 0.0019 to 0.1226 with different surfactants and corre-
structure and composition for magnetite. Note that the significance sponding xMS values ranged from 0.496 to 0.282. Such remarkable
of ı is consistent with that of xMS , of which both the values reflect variations demonstrated that surfactants indeed play an impor-
the defects of the magnetite in structure and composition to the tant role in regulating the crystal lattice defects and stoichiometry
extent of ideal crystalline. For former, the smaller the ı value, the of the nanoparticles synthesized by microemulsion method. Thus,
more ideal Fe3 O4 with less lattice defect will be obtained; and for according to the structural characteristics of surfactants, the sur-
later, the more close to 0.50 of xMS , the more perfect stoichiometric factant effects on the microstructure and composition of Fe3 O4
magnetite will be realized. nanoparticles, i.e. the extent of ideal crystalline, were discussed
as follows:
2 − SB /SA (a) Effect of headgroup charge: According to the results in
ı= (2)
6 + 5(SB /SA ) Table 3, the vacancy numbers of the samples synthesized by
680 T. Lu et al. / Colloids and Surfaces A: Physicochem. Eng. Aspects 436 (2013) 675–683

different surfactants followed SDS > Brij30 > DTAB, i.e. anionic hydrophobic tail [41]. That is to say, the slow relaxation time of
surfactant > nonionic surfactant > cationic surfactant. Also, the micelle formation-breakup which is associated with the lifetime of
magnetite stoichiometry implied the same trend, i.e. the order of the micelle, is strongly dependent on the hydrophobicity of the sur-
extent to near ideal Fe2+ content was DTAB > Brij30 > SDS. Both of factant; the longer of the hydrophobic tail, the longer of the micellar
the above results indicated that the lattice perfection of our Fe3 O4 lifetime is possessed [41,42]. Such long lifetime of reverse micelle
nanoparticles synthesized by microemulsion method decreased affects the process of nanodroplets’ motion-collision exchange and
from cationic surfactant to nonionic surfactant to anionic surfac- thus the growth rate of the nanoparticles. The lower growth rate
tant. Note that the above three surfactants are equal in length of of the crystallite is beneficial to improve the quality of the crys-
hydrophobic chain and the hydrophilic groups are relatively small tal, thus decreasing the lattice imperfection. Similar results were
which can be seen as one entity; therefore, the difference of the also observed in the systems of DEAB and CEAB (cetyltriethylam-
crystal defect is principally induced by the charge of the polar monium bromide) with different length of hydrophobic tail.
headgroup. It is known that the process of collision, fusion, separa- (c) Effect of headgroup size: It is interesting to find that the
tion and restructuring among the microemulsion nanodrops leads extent of crystal defect or the deviation of Fe3 O4 from stoichiom-
to the nucleation and subsequent growth of the product, result- etry was reduced when changing the conventional single-chain
ing in the formation of nanoparticles. Thus, the dynamic exchange cationic quaternary ammonium surfactant from trimethylammo-
process of the reactants in different water nanodomains plays an nium headgroup (DTAB) to triethylammonium group (DEAB) to
important role in controlling the quality of the crystallite. Based tributylammonium group (DBAB). That is to say, the bigger of the
on this viewpoint, the effect of headgroup charge can be rational- surfactant headgroup spacing is used, the more perfect of the Fe3 O4
ized in terms of the electrostatic interaction on the rate control crystallite can be obtained. Comparing with the above surfactants,
of iron ions diffusion and meeting with the OH− . In anionic sur- the mean area per headgroup increased with the surfactant chang-
factant of SDS microemulsion, the iron ions of Fe2+ and Fe3+ were ing from DTAB to DEAB to DBAB, resulting in the loose arrangement
attracted with the anionic headgroup SO4 − in the core of the reverse among the surfactant molecules in the reverse micelles. Although
micelles, hence there was a competition between SO4 − and OH− the increase of the mean area per headgroup is disadvantageous to
reacting with iron ions. Such an electrostatic interaction between the formation of reverse micelles according to the critical packing
surfactant and reactant would decrease the effective collisions for parameter theory, it is convenient to obtain reverse micelles with
material exchange and reaction, resulting in bad quality and perfec- flexible films. It is well known that not all droplets’ collisions are
tion of the products. Contrarily, the cationic surfactant DTAB is in effective for reactant exchange, and the film flexibility determines
favor of the reactants’ exchange which was accelerated by electro- the effective encounter rate factor of the collisions. For flexible film
static repulsion, thus bringing an improvement in the perfection like DBAB microemulsion, the effective encounter rate as well as the
of crystallite lattice. As for nonionic surfactant Brij30, there was microemulsion exchange characteristic time is faster than that of
no obvious electrostatic interaction between the headgroup and the rigid film [24]. Therefore, the increasing headgroups vary the
reactants, so the quality of the crystallite derived from Brij30 emul- fluidity of the interface and thus the kinetics of the intermicellar
sion was between that of cationic surfactant and anionic surfactant. exchange, which in turn ensures a more homogeneous reparti-
Similar arranging trend of cationic surfactant, nonionic surfactant tion of the reactants among different nanodroplets. This kinetic
and anionic surfactant were also observed in the other surfactants improvement favors the nucleation and growth of the nanocrys-
with bigger headgroup size, namely, DBAB (|ı| = 0.0378) < Triton tallite, and so does the formation of Fe3 O4 nanoparticles with less
X-100 (polyethylene ether, |ı| = 0.0787, Fig. 1 in Supplementary crystal defect and more close to stoichiometric magnetite.
material) < SDS (|ı| = 0.1226). Moreover, the above trend of cationic (d) New Gemini surfactant: As a new class of surfactants,
surfactant and anionic surfactant was also exhibited when the Gemini surfactants have attracted increasing attention in both
synthesized temperature was changed (detail in the latter part academic and industrial circles over the last years, owing to
of discussion on temperature effects), namely, DTAB (|ı| = 0.1317, their superior properties in comparison with those of conven-
50 ◦ C) < SDS (|ı| = 0.1546, 50 ◦ C) and DTAB (|ı| = 0.1268, 60 ◦ C) < SDS tional single-chain surfactants [43–47]. About Gemini surfactants,
(|ı| = 0.1385, 60 ◦ C), suggesting that the charge effect is the main there have the relevant reports on the application of tem-
factor for the generation of different vacancy numbers, i.e. crys- plate synthesis of various materials, such as gold particles and
tal defects. The changing trend of calculated values of magnetite organized mesoporous silica [48–50]. Here, a Gemini surfac-
stoichiometry (xMS ) was in good agreement with that of vacancy tant dimethylene-1,2-bis(dodecyldimethylammonium bromide),
numbers (ı). It should be mentioned that the ammonia was excess abbreviated as 12-2-12, has been devised as surfactant phase
in contrast to iron ions, therefore the effect of electrostatic interac- to synthesize the nanosized Fe3 O4 particles in water-in-oil
tion on reactants’ exchange can be ignored. microemulsion. It is interesting to find that high quality magnetite
(b) Effect of hydrophobic chain length: It is shown in Table 3 that with very little derivation from the stiochiometry was obtained
the crystal defect decreased with the increase of hydrophobic chain in the microemulsion of 12-2-12. The ı value which reflects the
length. When the hydrophobic chain of the surfactants changed defects of the sample in structure and composition to the extent of
from dodecyl (DTAB) to cetyl (CTAB) with the same headgroup, perfect crystalline was much smaller than that of conventional sur-
the absolute deviation parameter of synthesized Fe3 O4 decreased factants (Table 3) and also the magnetite stoichiometry was nearly
from 0.0858 to 0.0459, indicating the improvement of lattice per- close to ideal stoichiometric magnetite (x = 0.50), indicating that
fection. Based on critical packing parameter theory, the increase Gemini surfactant 12-2-12 plays positive effect on the nucleation
of the hydrophobic chain length produces a bigger critical pack- and growth of the nanocrystalline. It has been proved by Ulbricht
ing parameter p (p is defined as v/a0 lc , where v is the surfactant and Zana [51] that the rate constant for the entry of a Gemini surfac-
tail volume, lc is the tail length, and a0 is the equilibrium area per tant into its micelles, as compared with conventional surfactants,
molecule at the aggregate surface; 0 ≤ p ≤ 1/3 for sphere micelle, is slower than that for a diffusion-controlled process. Also the resi-
1/3 ≤ p ≤ 1/2 for cylinder micelle, 1/2 ≤ p ≤ 1 for bilayer structure, dence time of a Gemini surfactant in its micelles, i.e. the lifetime of
and p > 1 for reverse micelle) [40], which favors the formation Gemini surfactant micelles is much longer than for the correspond-
of reverse micelles. On the other hand, it has been proved that ing conventional surfactants [51]. The long lifetime of the Gemini
the lifetime of the ionic surfactant micelles in aqueous solution, surfactant micelles prolongs the whole process of motion-collision
i.e. the characteristic time of micellization-dissolution equilib- exchange ( ex ) between different nanodomains, even equal or
rium, shows a noteworthy increase with increasing length of the greater to the characteristic chemical reaction time ( r ). Thus, the
T. Lu et al. / Colloids and Surfaces A: Physicochem. Eng. Aspects 436 (2013) 675–683 681

Table 4
Average particle diameters (D) of Fe3 O4 nanoparticles synthesized at different tem-
perature calculated through TEM observation.

Surfactants D (nm)

50 ◦ C 60 ◦ C 70 ◦ C

SDS 12.8 12.8 13.0


DTAB 14.0 13.9 14.2
CTAB 13.9 14.1 14.2
DEAB 13.9 14.1 14.2

slow process of the chemical reaction, which is restricted by the


process of motion-collision exchange, favors the nucleation and
growth of the Fe3 O4 nanocrystalline, showing better crystal per-
fection and stoichiometry. Moreover, owing to much lower critical
micelle concentration (cmc) and relatively small aggregation num-
ber [29,52], the micelle concentration of Gemini surfactant was
larger than those of conventional surfactant, thus increasing the
probability of effective collisions for material exchange, which is
Fig. 5. Variation of vacancy parameter (ı) against various surfactants as a function
also beneficial to the formation of Fe3 O4 nanoparticles. of temperature. The inset is the absolute value of ı changed from 50 ◦ C to 70 ◦ C.

3.3. Effects of temperature on particles size and crystal defect


the data of Table 3 both in the text and Supplementary material,
Since the same synthesis procedure was adopted, namely, same the variation of vacancy parameter (ı) against temperature as a
reactant concentration, same reaction temperature and same reac- function of various surfactants is shown in Fig. 5. It can be noted
tion pressure, no different kinetic parameters can be obtained from that with the increase of temperature from 50 ◦ C to 70 ◦ C, the crys-
the viewpoint of reaction kinetics. However, from the viewpoint tal defect decreased, i.e. the higher the synthesized temperature,
of crystal formation kinetics, the different surfactants indeed play the more perfect crystallite would be obtained. This fact can be
an important role in crystal nucleation and growth, and conse- explained from the influence of temperature on the microemul-
quently affect the perfection of crystallite lattice and stoichiometry. sion dynamics. With the increase of temperature, the viscosity of
To further understand the surfactant effect on the microstructures the microemulsion systems decreases, and the Brownian motion
of Fe3 O4 nanoparticles, kinetics correlative experiments were car- becomes more active, thus increasing the effective collision and
ried out, involving the temperature effects on the particle size and favoring the nucleation and growth of nanoparticles. To further
crystal defect. determine the proportion of the surfactant structural factors on
Table 4 shows the average particle diameters (D) of Fe3 O4 the crystal defect during the variation of temperature, a histogram
nanoparticles synthesized at different temperature calculated of the variation of ı value vs. different surfactants against the
through TEM observation. Similar to the samples synthesized at increase of temperature from 50 ◦ C to 70 ◦ C, is shown in the inset of
70 ◦ C (Fig. 2), XRD results showed that single-phase of spinel-type Fig. 5. Obviously, among the various surfactant structural factors,
Fe3 O4 can also be obtained when the synthesized temperature was including headgroup charge (SDS vs. DTAB), hydrophobic chain
decreased to 60 ◦ C and 50 ◦ C (Fig. 2 in Supplementary material). length (DTAB vs. CTAB) and headgroup size (DTAB vs. DEAB), the
TEM images for the above Fe3 O4 samples derived from microemul- length of the hydrophobic tail exhibited the most distinct change
sions of various surfactants at 50 ◦ C and 60 ◦ C are shown in Fig. against the temperature variation, while the rest almost took the
3 in Supplementary material. It can be seen that roughly spheri- same extent.
cal particles with a few rod-like aggregates were exhibited. Also,
statistical size distribution results showed no obvious variation of
particle size can be observed during the synthesized temperature
regulated from 50 ◦ C to 70 ◦ C (Table 4), demonstrating that tem-
perature variation does not have a remarkable influence on particle
size. In other words, in spite of various surfactants with different
headgroup charge (SDS vs. DTAB), hydrophobic chain length (DTAB
vs. CTAB) and headgroup size (DTAB vs. DEAB), no obvious temper-
ature effects on the particle size can be observed by microemulsion
method.
However, further investigation indicates that the temperature
plays an important role in regulating the vacancy number and
stoichiometry of Fe3 O4 nanoparticles. Room temperature 57 Fe
Mössbauer spectra of the as-prepared Fe3 O4 nanoparticles at 50 ◦ C
and 60 ◦ C (Fig. 4 in Supplementary material) still showed two
six-line sub-spectra fitting assigned to tetrahedral (A) sites and
octahedral (B) sites of Fe cations, which was similar to those of
the corresponding samples synthesized at 70 ◦ C. Also, the IS and QS
parameters derived from Mössbauer spectra exhibited no obvious
change with the decrease of temperature, except for a remarkable
change of the relative peak areas of A sites and B sites (Table 2
Fig. 6. Magnetization curves for Fe3 O4 nanoparticles synthesized in microemulsions
in Supplementary material), implying the variation of the devia- of various surfactants when tested at room temperature: (a) SDS, (b) Brij30, (c) DTAB,
tion parameters (ı) from stoichiometry of magnetite. Combined (d) DEAB, (e) CTAB, (f) 12-2-12. The inset is the local image magnified.
682 T. Lu et al. / Colloids and Surfaces A: Physicochem. Eng. Aspects 436 (2013) 675–683

3.4. Vibrating sample magnetometry (VSM) measurements [5] S.X. Zhang, X.L. Zhao, H.Y. Niu, Y.L. Shi, Y.Q. Cai, G.B. Jiang, Superparamag-
netic Fe3 O4 nanoparticles as catalysts for the catalytic oxidation of phenolic
and aniline compounds, J. Hazard. Mater. 167 (2009) 560–566.
The magnetic properties of the Fe3 O4 nanoparticles were mea- [6] B. Sreedhar, A.S. Kumar, P.S. Reddy, Magnetically separable Fe3 O4 nanoparti-
sured, as shown in Fig. 6. As seen, the values of saturation cles: an efficient catalyst for the synthesis of propargylamines, Tetrahedron
magnetization (MS ) of most samples were approximately equal Lett. 51 (2010) 1891–1895.
[7] P. Moodley, F.J.E. Scheijen, J.W. Niemantsverdriet, P.C. Thüne, Iron oxide
regardless of vacancy numbers (ı), however, with further increase nanoparticles on flat oxidic surfaces—Introducing a new model catalyst for
of vacancy number (|ı| = 0.1226 for SDS), MS value decreased. It Fischer-Tropsch catalysis, Catal. Today 154 (2010) 142–148.
is indicated that the crystal perfection Fe3 O4 nanoparticles influ- [8] Y.K. Kim, M. Oliveria, Magnetic-properties and texture of sputtered Fe/Fe3 O4
multilayer films, IEEE Trans. Magn. 30 (1994) 1316–1323.
ences their saturation magnetization; on the whole, the existence
[9] L.Q. Yu, L.J. Zheng, J.X. Yang, Study of preparation and properties on magne-
of crystal defect decreases saturation magnetization of Fe3 O4 tization and stability for ferromagnetic fluids, Mater. Chem. Phys. 66 (2000)
nanoparticles. Moreover, the presence of coercivity and remanence 6–9.
[10] W.M. Zhang, X.L. Wu, J.S. Hu, Y.G. Guo, L.J. Wan, Carbon coated Fe3 O4 nanospin-
indicates that the Fe3 O4 nanoparticles synthesized did not possess
dles as a superior anode material for lithium-ion batteries, Adv. Funct. Mater.
superparamagnetic behavior, which is consistent with the results 18 (2008) 3941–3946.
of Mössbauer characterization. [11] I. Bica, Nanoparticle production by plasma, Mater. Sci. Eng. B 68 (1999) 5–9.
[12] K.M. Bratlie, H. Lee, K. Komvopoulos, P. Yang, G.A. Somorjai, Platinum nanopar-
ticle shape effects on benzene hydrogenation selectivity, Nano Lett. 7 (2007)
4. Perspective and conclusions 3097–3101.
[13] D. Amara, I. Felner, I. Nowik, S. Margel, Synthesis and characterization of Fe
and Fe3 O4 nanoparticles by thermal decomposition of triiron dodecacarbonyl,
Water-in-oil microemulsions containing appropriate amount of Colloids Surf. A: Physicochem. Eng. Aspects 339 (2009) 106–110.
n-heptane as the oil phase, n-hexanol as the co-surfactant phase, [14] J.Y. Chen, B. Lim, E.P. Lee, Y.N. Xia, Shape-controlled synthesis of platinum
nanocrystals for catalytic and electrocatalytic applications, Nano Today 4
and various surfactants as the surfactant phase respectively, have (2009) 81–95.
been applied to synthesize Fe3 O4 nanoparticles with the average [15] G.F. Goya, Handling the particle size and distribution of Fe3 O4 nanoparticles
particle size of <16 nm. The headgroup charge, hydrophobic chain through ball milling, Solid State Commun. 130 (2004) 783–787.
[16] Z.H. Zhou, J. Wang, X. Liu, H.S.O. Chan, Synthesis of Fe3 O4 nanoparticles from
length, headgroup size of conventional single-chain surfactants, emulsions, J. Mater. Chem. 11 (2001) 1704–1709.
and one Gemini surfactant have obvious effects on the microstruc- [17] R. Fan, X.H. Chen, Z. Gui, L. Liu, Z.Y. Chen, A new simple hydrothermal prepara-
ture crystallite characteristics and stoichimetry of the above Fe3 O4 tion of nanocrystalline magnetite Fe3 O4 , Mater. Res. Bull. 36 (2001) 497–502.
[18] Y.S. Kang, S. Risbud, J.F. Rabolt, P. Stroeve, Synthesis and characterization of
nanoparticles. There is a variation in vacancy parameter and sto-
nanometer-size Fe3 O4 and gamma-Fe2 O3 particles, Chem. Mater. 8 (1996)
ichimetry against various surfactant-derived Fe3 O4 particles by 2209–2211.
microemulsion method. The different surfactants act the dynamic [19] H.S. Lee, W.C. Lee, T.J. Furubayashi, A comparison of coprecipitation with
microemulsion methods in the preparation of magnetite, J. Appl. Phys. 85
exchange process of the reactants, film flexibility of the reverse
(1999) 5231–5233.
micelles and the lifetime of a surfactant molecule in the micelle, all [20] F. Grasset, N. Labhsetwar, D. Li, D.C. Park, N. Saito, H. Haneda, O. Cador, T. Roisnel,
of which play critical roles in controlling the nucleation and growth S. Mornet, E. Duguet, J. Portier, J. Etourneau, Synthesis and magnetic charac-
of the nanocrystalline, thus affecting the lattice defects of the sam- terization of zinc ferrite nanoparticles with different environments: powder,
colloidal solution, and zinc ferrite-silica core-shell nanoparticles, Langmuir 18
ple in structure and composition to the extent of perfect crystalline (2002) 8209–8216.
and briningg the deviation of magnetite from stoichiometry. It is [21] R. Jain, A. Mehra, Monte Carlo models for nanoparticle formation in two
also found that the synthesized temperature of samples plays an microemulsion systems, Langmuir 20 (2004) 6507–6513.
[22] M. Sanchez-Dominguez, M. Boutonnet, C. Solans, A novel approach to metal and
important role in regulating the crystal defects owing to the crystal metal oxide nanoparticle synthesis: the oil-in-water microemulsion reaction
formation kinetics, and the hydrophobic chain length of the surfac- method, J. Nanopart. Res. 11 (2009) 1823–1829.
tants shows the most remarkable influence on the crystal defect [23] J.H. Schulman, W. Stoeckenius, L.M. Prince, Mechanism of formation and struc-
ture of micro emulsions by electron microscopy, J. Phys. Chem. 63 (1959)
and stoichiometry against the temperature variation. Moreover, 1677–1680.
magnetic properties studies show that the worse the crystal defect [24] M.A. López-Quintela, C. Tojo, M.C. Blanco, L. García Rio, J.R. Leis, Microemulsion
of Fe3 O4 nanoparticles, the smaller the saturation magnetization dynamics and reactions in microemulsions, Curr. Opin. Colloid Interface Sci. 9
(2004) 264–278.
is. We hope this work will provide a powerful help in choosing a
[25] M. Boutonnet, S. Lögdberg, E.E. Svensson, Recent developments in the appli-
surfactant to synthesize nanoparticles by emulsion method. cation of nanoparticles prepared from w/o microemulsions in heterogeneous
catalysis, Curr. Opin. Colloid Interface Sci. 13 (2008) 270–286.
[26] T. Aubert, F. Grasset, S. Mornet, E. Duguet, O. Cador, S. Cordier, Y. Molard, V.
Acknowledgments Demange, M. Mortier, H. Haneda, Functional silica nanoparticles synthesized
by water-in-oil microemulsion processes, J. Colloid Interface Sci. 341 (2010)
201–208.
Financial support by the National Natural Science Foundation of [27] P. Kumar, K.L. Mittal, Handbook of Microemulsion Science and Technology,
China (21003124, 21076211) is gratefully acknowledged. Marcel Dekker, Inc, New York, 1999.
[28] S. Santra, R. Tapec, N. Theodoropoulou, J. Dobson, A. Hebard, W.H. Tan, Synthesis
and characterization of silica-coated iron oxide nanoparticles in microemul-
Appendix A. Supplementary data sion: the effect of nonionic surfactants, Langmuir 17 (2001) 2900–2906.
[29] R. Zana, M. Benrraou, R. Rueff, Alkanediyl-␣,(-bis(dimethylalkylammonium
bromide) surfactants: 1. Effect of the spacer chain length on the critical micelle
Supplementary data associated with this article can be concentration and micelle ionization degree, Langmuir 7 (1991) 1072–1075.
found, in the online version, at http://dx.doi.org/10.1016/j.colsurfa. [30] J.A. López Pérez, M.A. López Quintela, J. Mira, J. Rivas, S.W. Charles, Advances
in the preparation of magnetic nanoparticles by the microemulsion method, J.
2013.08.004. Phys. Chem. B 101 (1997) 8045–8047.
[31] R. Mcgehee, J. Renault, Use of standard deviation of X-ray-diffraction lines as
a measure of broadening in Scherrer equation: curve fitting method, J. Appl.
References Crystallogr. 50 (1972) 365–369.
[32] K. Voleník, M. Seberíni, J. Neid, A Mössbauer and X-ray-diffraction study of
[1] K. Cornelis, C.S. Hurlburt, Manual of Mineralogy, Wiley, New York, 1977. nonstoichiometry in magnetite, Czech. J. Phys. B 25 (1975) 1063–1071.
[2] A.K. Gupta, M. Gupta, Synthesis and surface engineering of iron oxide nanopar- [33] A.F. Lehlooh, S.H. Mahmood, Mössbauer-spectroscopy of Fe3 O4 ultrafine parti-
ticles for biomedical applications, Biomaterials 26 (2005) 3995–4021. cles, J. Magn. Magn. Mater. 151 (1995) 163–166.
[3] Y. Zhang, J. Zhang, Surface modification of monodisperse magnetite nanopar- [34] X.D. Wang, X.Q. Zhao, J.Y. Shen, X.Y. Sun, T. Zhang, L.W. Lin, A Mössbauer study
ticles for improved intracellular uptake to breast cancer cells, Colloid Interface of In-Fe2 O3 /HZSM-5 catalysts for the selective catalytic reduction of NO by
Sci. 283 (2005) 352–357. methane, Phys. Chem. Chem. Phys. 4 (2002) 2846–2851.
[4] L.Z. Gao, J. Zhuang, L. Nie, J.B. Zhang, Y. Zhang, N. Gu, T.H. Wang, J. Feng, D.L. Yang, [35] E. Schmidbauer, M. Keller, Magnetic hysteresis properties, Mössbauer spectra
S. Perrett, X. Yan, Intrinsic peroxidase-like activity of ferromagnetic nanopar- and structural data of spherical 250 nm particles of solid solutions Fe3 O4 -␥-
ticles, Nat. Nanotechnol. 2 (2007) 577–583. Fe2 O3 , J. Magn. Magn. Mater. 297 (2006) 107–117.
T. Lu et al. / Colloids and Surfaces A: Physicochem. Eng. Aspects 436 (2013) 675–683 683

[36] J. Yin, J.H. Wang, T. Zhang, X.D. Wang, Novel alumina-supported PtFe alloy [44] M.J. Rosen, Geminis: a new generation of surfactants, Chem. Tech. 23 (1993)
nanoparticles for preferential oxidation of carbon monoxide in hydrogen, Catal. 30–33.
Lett. 125 (2008) 76–82. [45] F.M. Menger, J.S. Keiper, Geminis surfactants, Angew. Chem. Int. Ed. 39 (2000)
[37] E. Murad, J. Cashion, Mössbauer Spectroscopy of Environmental Materials 1906–1920.
and their Industrial Utilization, Kluwer Academic Publishers, USA, 2004, pp. [46] R. Zana, Dimeric and oligomeric surfactants. Behavior at interfaces and in aque-
159–188. ous solution: a review, Adv. Colloid Interface Sci. 97 (2002) 205–253.
[38] C.A. Gorski, M.M. Scherer, Determination of nanoparticulate magnetite stoi- [47] T. Lu, F. Han, G.R. Mao, G.F. Lin, J.B. Huang, X. Huang, Y.L. Wang, H.L. Fu, Effect of
chiometry by Mössbauer spectroscopy, acidic dissolution, and powder X-ray hydrocarbon parts of the polar headgroup on surfactant aggregates in gemini
diffraction: a critical review, Am. Mineral. 95 (2010) 1017–1026. and bola surfactant solutions, Langmuir 23 (2007) 2932–2936.
[39] U. Colombo, F. Gazzarrini, G. Lanzavecchia, Mechanisms of iron oxides [48] Q. Huo, R. Leon, P.M. Petroff, G.D. Stucky, Mesostructure design with gemini sur-
reduction at temperatures below 400 ◦ C, Mater. Sci. Eng. 2 (1967) factants: supercage formation in a three-dimensional hexagonal array, Nature
125–135. 268 (1995) 1324–1327.
[40] (a) J.N. Israelachvili, D.J. Mitchell, B.W. Ninham, Theory of self-assembly of [49] S. Sato, K. Toda, S. Oniki, Kinetic study on the formation of colloidal gold in the
hydrocarbon amphiphiles into micelles and bilayers, J. Chem. Soc. Faraday presence of acetylenic glycol nonionic surfactant, J. Colloid Interface Sci. 218
Trans. II 72 (1976) 1525–1568; (1999) 504–510.
(b) J.N. Israelachvili, D.J. Mitchell, B.W. Ninham, Theory of self-assembly of lipid [50] S.H. Han, J. Xu, W.G. Hou, X.M. Yu, Y.S. Wang, Synthesis of high-quality
bilayers and vesicles, Biochim. Biophys. Acta 470 (1977) 185–201; MCM-48 mesoporous silica using gemini surfactant dimethylene-1,2-
(c) J.N. Israelachvili, S. Marčelja, R.G. Horn, Physical principles of membrane bis(dodecyldimethylammonium bromide), J. Phys. Chem. B 108 (2004)
organization, Rev. Biophys. 13 (1980) 121–200. 15043–15048.
[41] E.A.G. Aniansson, S.N. Wall, M. Almgren, H. Hoffmann, I. Klelmann, W. Wlbricht, [51] W. Ulbricht, R. Zana, Alkanediyl-␣,␻-bis(dimethylalkylammonium bromide)
R. Zana, J. Lang, C. Tondre, Theory of kinetics of micellar equilibria and quan- surfactants: part 8. Pressure-jump study of the kinetics of micellar equilibria
titative interpretation of chemical relaxation studies of micellar solutions of in aqueous solutions of alkanediyl-␣,␻-bis(dimethyldodecylammonium bro-
ionic surfactants, J. Phys. Chem. 80 (1976) 905–922. mide) surfactants, Colloids Surf. A: Physicochem. Eng. Aspects 183–185 (2001)
[42] J. Lang, R. Zana, in: R. Zana (Ed.), Surfactant Solutions. New Methods of Inves- 487–494.
tigation, Marcel Dekker, New York, 1987, p. 405. [52] D. Danino, Y. Talmon, R. Zana, Alkanediyl-␣,␻-bis(dimethylalkylammonium
[43] F.M. Menger, C.A. Littau, Gemini surfactants: synthesis and properties, J. Am. bromide) surfactants (dimeric surfactants). 5. Aggregation and microstructure
Chem. Soc. 113 (1991) 1451–1452. in aqueous solutions, Langmuir 11 (1995) 1448–1456.

Potrebbero piacerti anche