Sei sulla pagina 1di 7

FULL PAPER

Emergent Nanostructures: Water-Induced Mesoscale


Transformation of Surfactant-Stabilized Amorphous Calcium
Carbonate Nanoparticles in Reverse Microemulsions**
By Mei Li and Stephen Mann*

The water-induced crystallization of alkylbenzenesulfonate-coated amorphous calcium carbonate (ACC) nanoparticles in


water-in-isooctane sodium bis(2-ethylhexyl)sulfosuccinate (NaAOT) microemulsions at a [H2O]/[NaAOT] molar ratio (w) of
10:1 produces a range of organized hybrid surfactant±vaterite nanostructures depending on the water droplet/ACC nanoparti-
cle number ratio, n = [H2O]/[CaCO3]. The crystalline nanostructures develop within primary aggregates of the surfactant-stabi-
lized ACC nanoparticles by in situ mesoscale transformation, which is mediated by the extent of coupling at the surfactant±in-
organic interface. Strong coupling in the presence of low amounts of water (n = 34) gives monodisperse spheroidal aggregates
of densely packed 5 nm diameter surfactant-coated vaterite nanoparticles, whereas weak interactions at n = 3400 produce dis-
crete vaterite nanoparticles, 130 nm in size. Significantly, intermediate levels of coupling produce anisotropic nanostructures
such as spindle-shaped aggregates of 18 nm sized surfactant-coated vaterite nanoparticles (n = 170) and high-aspect-ratio bun-
dles of co-aligned 10 nm wide twisted vaterite nanofilaments (n = 340). Adding excess aqueous CO32± to the microemulsion
droplets inhibits the growth of the nanofilaments, whereas excess Ca2+ has no effect. The results show that the transformation
pathways are determined by the extent of water penetration into the ACC cores and electrostatic interactions at the mineral±
surfactant interface, and indicate that complex hybrid nanostructures can be assembled in situ when these processes are
coupled synergistically at the mesoscopic level. Such observations could be of generic importance in nanochemistry and bio-
mineralization.

1. Introduction bution because many biogenic minerals are inorganic±organic


nanostructures with long range organization and complex mor-
The synthesis of organized structures with dimensions of phological form. These materials arise by vectorial shaping and
1±100 nm is a dominant theme of nanochemistry. Although patterning of vesicles or polymeric frameworks, and reflect a
many approaches in nanochemistry seek to mimic the informa- synergy between the force fields of inorganic precipitation and
tion processing and responsive sensing capabilities of biological biological organization.[13]
nanostructures, they often lack the inherent materials-building The complex interplay between crystallization processes and
properties of organisms, which are essential if nanostructures environmental constraints has been investigated recently by
are to be organized across many length scales and used as func- studying the structural evolution of inorganic solids synthesized
tional materials within integrated systems. For this reason a in organized reaction media. For example, macroscopic struc-
number of strategies have been developed to physically pattern tures consisting of surfactant±inorganic nanofilaments orga-
and control the long-range organization and assembly of nano- nized into twisted bundles, cones, and helicoids of BaSO4,
structured phases. These include the combination of chemical BaCrO4 or Ba phosphotungstate,[14] BaCO3,[15] or CaSO4,[16]
and microfabrication methods,[1,2] and controlled deposition of have been synthesized in water-in-oil microemulsions. Similarly,
nanoparticle-based superlattices using solvent evaporation,[3,4] nested polymer±inorganic colloids containing neuron-like tan-
molecular cross-linking[5,6] or programmed recognition.[7±9] Al- gles of calcium phosphate nanofilaments have been prepared
though promising, these approaches involve sequential pro- within aggregates of a partially alkylated poly(methacrylic
cessing, and alternative methods based on the in situ coupling acid)±poly(ethylene glycol) (PMAA±PEG) block copoly-
of synthesis and self-assembly should offer significant advan- mer.[17] A key feature of these systems lies in their interactive
tages.[10] In this regard, biomimetic approaches inspired by the nature, which determines the time-dependent coupling between
study of biomineralization[11,12] can make an important contri- inorganic growth and associated adaptations in the surrounding
reaction environment. As a consequence, a series of co-opera-
tive feedback loops are established, and the systems are notable
± for their emergent properties. Significantly, similar observations
[*] Prof. S. Mann, Dr. M. Li have been reported for inorganic crystallization in biopoly-
School of Chemistry, University of Bristol mer[18] or silica[19] gels, as well as in aqueous solutions of anionic
Bristol BS8 1TS (UK)
polyelectrolytes,[20] suggesting that a generic mechanism could
E-mail: s.mann@bris.ac.uk
[**] We thank Dr. Julian Eastoe for the gift of surfactant-stabilized amorphous
be responsible for the emergence of hybrid nanostructures in di-
calcium carbonate particles, and EPSRC for financial support to ML. verse synthesis regimes.

Adv. Funct. Mater. 2002, 12, No. 11±12, December Ó 2002 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim 1616-301X/02/1112-0773 $ 17.50+.50/0 773
S. Mann, M. Li/Emergent Nanostructures: Mesoscale Transformation of CaCO3 Nanoparticles
FULL PAPER

A common feature of many of the above examples is that appearance of a white precipitate at the bottom of the test-
nucleation and initial growth of the nanostructures occur with- tube. At [H2O]/[CaCO3] = 34, scanning electron microscopy
in aggregates of surfactant- or polymer-coated inorganic nano- (SEM) showed that the precipitate consisted of spheroidal par-
particles. Significantly, the nanoparticles are amorphous and ticles with a highly monodisperse distribution and a mean size
often only 5 nm or so in diameter due to surface interactions of 180 nm (r = 6 nm, n = 50), which comprised densely packed
with the surrounding surfactant/polymer molecules that stabi- disordered clusters of smaller spherical nanoparticles (Fig. 1a).
lize the disordered structure, restrict particle growth, and in- Corresponding transmission electron microscopy (TEM)
duce aggregation into the larger colloidal clusters. With time, images indicated that the primary nanoparticles were ca. 5 nm
however, the amorphous nanoparticles slowly transform within
the aggregates to their crystalline counterparts, and because A
this process is strongly coupled with the surface-adsorbed mol-
ecules, the mineral phase becomes structurally reorganized and
morphologically modulated across a range of length scales.
It follows therefore from the above hypothesis that meso-
scale transformations are of pivotal importance, and for this
reason we now report a systematic study of the controlled crys-
tallization of surfactant-coated amorphous calcium carbonate
(ACC) nanoparticles. The ACC nanoparticles are used com-
mercially as a pH buffer in engine oils, and consist of a 5 nm
diameter inorganic core, surrounded by a surfactant shell com-
prising a mixture of calcium alkylbenzenesulfonates with a dis-
500 nm
tribution of alkyl chain lengths centred on n-C24.[21,22] By using
preformed nanoparticles rather than reaction solutions we are
B
able to decouple the primary nucleation events from the subse-
quent phase transformation processes. Moreover, because
ACC rapidly transforms in the presence of water to crystalline
polymorphs such as aragonite and calcite,[23±25] we can achieve
a high level of control over this process by introducing very
low amounts of water into oil-based suspensions of the ACC
nanoparticles. For this, we use water-in-isooctane reverse
microemulsions prepared from the surfactant, NaAOT (sodium
bis(2-ethylhexyl)sulfosuccinate), with a typical [H2O]/[Na-
AOT] molar ratio (w value) of 10. We show that the addition
of water droplets induces aggregation of the surfactant-coated 200 nm
ACC nanoparticles, followed by in situ transformation at the
Fig. 1. a) SEM and b) TEM images of spheroidal aggregates of surfactant-stabi-
mesoscopic level to produce hybrid nanostructures containing lized vaterite nanoparticles produced by phase transformation of ACC in
the crystalline calcium carbonate polymorph, vaterite. Changes NaAOT reverse microemulsions ([H2O]/[NaAOT]/[CaCO3] = 34:3.4:1. Inset in
in the number of water droplets per ACC nanoparticle has a (b) shows electron diffraction pattern obtained from the aggregates; d spacings,
4.24 Š (002), 3.57 Š (100), 3.28 Š (101), 2.72 Š (102), 2.06 Š (110), 1.83 Š
marked influence on the growth and higher-order assembly of (104) and 1.64 Š (202).
the vaterite nanocrystals, such that organized bundles of surfac-
tant±vaterite nanofilaments can be produced under specific
conditions. in size (Fig. 1b), and energy dispersive X-ray (EDX) analysis
showed the presence of calcium (3.7, 4.0 keV), sulfur
(2.3 keV), and sodium (1.0 keV) in the aggregates (Fig. 2a).
2. Results and Discussion Significantly, selected area electron diffraction studies showed
that the constituent nanoparticles were crystalline, with diffrac-
2.1. Influence of Water Droplet/ACC Number Ratio at w = 10 tion rings consistent with the vaterite crystal structure (Fig. 1b,
inset). The presence of vaterite was confirmed by powder
A series of experiments were undertaken in which the num- X-ray diffraction analysis (Fig. 2b), and FTIR spectroscopy,
ber of water droplets per ACC nanoparticleÐas defined by the which showed absorption bands for CO32± at 878, 1088, and
[H2O]/[CaCO3] molar ratioÐpresent in microemulsions 1450 cm±1. FTIR spectra also showed bands corresponding to
prepared at constant surfactant concentration ([NaAOT = organic functionalities, such as ester C=O (1737 cm±1), ester
0.1 M) and w value ([H2O]/[NaAOT] = 10) was systematically C=O, and S=O (1218 cm±1), C=C aromatic (1493 cm±1), C±H
changed. In general, adding dry samples of the surfactant-stabi- methyl (2960, 2874 cm±1), C±H methylene (2931, 2859 cm±1),
lized ACC nanoparticles to the NaAOT microemulsions pro- and O=S=O (1180 cm±1) (Fig. 2c), indicating that a mixture of
duced transparent solutions that became slightly turbid within alkylbenzenesulfonate and AOT surfactants were present in
30 min and cloudy after two hours, followed by the gradual the nanoparticle aggregates.

774 Ó 2002 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim 1616-301X/02/1112-0774 $ 17.50+.50/0 Adv. Funct. Mater. 2002, 12, No. 11±12, December
S. Mann, M. Li/Emergent Nanostructures: Mesoscale Transformation of CaCO3 Nanoparticles

FULL PAPER
a) cps phobic aggregate interior, where each ACC nanoparticle acts
50 Ca as a single nucleation centre. The nuclei remain spatially sepa-
rated, presumably due to the low level of water penetration
40 Cu and concomitant strong surface anchoring of the surfactant
30
molecules, such that growth of each vaterite crystallite is re-
stricted to the size of the ACC core (ca. 5 nm).
20 The presence of NaAOT in the vaterite aggregates suggests
that the microemulsion water droplets play an important role
10 in mediating the aggregation and crystallization processes. This
Ca Cu
Na S
was confirmed by increasing the number of water droplets per
0
0 5 10 ACC nanoparticle through reductions in the amount of added
Energy(keV) ACC. For example, at [H2O]/[CaCO3] = 170, no spheroidal
b) aggregates were observed; instead, spindle-shaped aggregates,
(101) (102)
280 nm (r = 32 nm, n = 30) and 90 nm (r = 3, n = 30) in mean
200
length and width, respectively, were precipitated (Fig. 3). Com-
(110) (104) pared with the spheroidal aggregates, the spindle-shaped clus-
Lin (Counts)

150 (100)
ters consisted of vaterite nanoparticles that were significantly
100

(202)
(002)
50

0
20 30 40 50 60
2-Theta - Scale
c)
80

60
%T

40

20

0
4000 3500 3000 2500 2000 1500 1000 500

cm-1
100 nm

Fig. 2. Characterization of surfactant-vaterite spheroidal aggregates. a) EDX Fig. 3. TEM image of spindle-shaped aggregates of vaterite nanoparticles pro-
analysis showing the presence of Ca, S and Na (Cu is from the TEM grid). duced by phase transformation of surfactant-stabilized ACC in NaAOT reverse
b) Powder X-ray diffraction profile showing {hkl} vaterite reflections (space microemulsions at [H2O]/[NaAOT]/[CaCO3] = 170:17:1. Inset shows correspond-
group, P63/mmc; a = 7.147 Š, c = 16.917 Š). c) FTIR spectrum (see text for ing vaterite powder electron diffraction pattern; d spacings, 4.24 Š (002), 3.57 Š
details). (100), 3.28 Š (101), 2.72 Š (102), 2.11 Š (004).

The above data indicate that crystallization of the vaterite larger (ca. 18 nm). Moreover, the nanoparticles were often
nanoparticles involves aggregation and transformation of the slightly oblate in shape and partially organized into disordered
surfactant-stabilized ACC precursors, without significant chains that were aligned approximately parallel to the spindle
change in the size of the inorganic cores. TEM studies on the long axis.
time-dependent evolution of the vaterite nanostructures Increasing the [H2O]/[CaCO3] molar ratio to 340 produced
showed that the onset of aggregation occurred within 30 min high aspect ratio aggregates in the form of elongated fibrous
of mixing the reactants, after which the aggregates increased bundles that were 490 nm (r = 15 nm, n = 30) and 53 nm (r = 5,
progressively in size over a period of 2 h. Significantly, the n = 30) in length and width, respectively, and which consisted of
aggregates were initially amorphous when studied by electron co-aligned vaterite nanofilaments (Fig. 4a). The individual
diffraction (after 40 min), indicating that vaterite nucleation nanofilaments were approximately 100 nm ” 10 nm in dimen-
occurs in situ within preformed clusters of the surfactant-stabi- sion, and although the length was somewhat variable, the width
lized ACC nanoparticles. The ratio of water droplets to nano- was remarkably uniform. Surprisingly, many of the nanofila-
particles is relatively low under these conditions, so the rate of ments were not straight but exhibited small regular twists along
particle coalescence, which was increased in the presence of their length, suggesting a helical ribbon nanoform. Moreover,
small amounts of water, exceeds the rate of water penetration although the filaments appeared to be entwined within individ-
into the cores of the ACC-surfactant nanoparticles. Vaterite ual bundles, they were not in direct contact but spaced at a
crystallization therefore occurs slowly within the highly hydro- distance of a few nanometers. The presence of NaAOT and

Adv. Funct. Mater. 2002, 12, No. 11±12, December Ó 2002 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim 1616-301X/02/1112-0775 $ 17.50+.50/0 775
S. Mann, M. Li/Emergent Nanostructures: Mesoscale Transformation of CaCO3 Nanoparticles
FULL PAPER

effectively with interparticle aggregation. This ac-


a
counts for the progressively larger nanoparticles
and lower nuclearity of the aggregates formed at
five- and tenfold increases in the [H2O]/[CaCO3]
molar ratio. Moreover, at w = 10 these molar
ratios are associated with the onset of shape
anisotropy in the crystal growth and aggregation
processes, suggesting a surfactant-mediated crys-
tallization pathway at intermediate levels of
water penetration. In this regard, TEM studies of
the early growth stages (between 10 and 40 mins
after mixing) showed a predominance of rod-
shaped primary aggregates, suggesting that vater-
ite crystallization and ACC nanoparticle aggrega-
tion were concurrent processes.
Reducing the concentrations of NaAOT, H2O,
and CaCO3 tenfold, but maintaining the molar
100nm ratios as in the above experiments, had no signifi-
cant effect on the morphology of the nanostruc-
50nm tures, implying that the molar ratios rather than
absolute concentrations were responsible for the
c variety of structures observed. Other experiments
indicated that changes in the pH of the water
droplets or temperature did not significantly alter
the morphology of the nanostructures. However,
the presence of aqueous CO32± or Ca2+ ions in the
microemulsion droplets had a significant effect
on the formation of the nanofilament architec-
b c
ture. For example, when the [CO32±]/[Ca2+] molar
Fig. 4. a) TEM image of co-aligned bundles of vaterite nanofilaments prepared by phase transfor- ratio was increased from 1:1 (above experiments)
mation of surfactant-stabilized ACC nanoparticles in NaAOT reverse microemulsions at [H2O]/
to 1.6:1 for [H2O]/[CaCO3] = 340, TEM studies
[NaAOT]/[CaCO3] = 340:34:1. Note the twisted-ribbon morphology. Inset shows corresponding
vaterite powder electron diffraction pattern; d-spacings, 4.26 Š (002), 3.59 Š (100), 3.26 Š (101), showed no evidence of nanofilament bundles;
2.73 Š (102), 2.09 Š (110), and 1.87 Š (112). b) single vaterite bundle with two nanofilaments instead, discrete irregular prismatic nanoparticles
showing direction of crystallographic c-axis determined by electron diffraction analysis as shown
in (c). c) single crystal electron diffraction pattern recorded from nanofilaments shown in (b). The
of vaterite (mean length and width, 190 nm
pattern corresponds to a view down the [010] zone of vaterite and shows the directions of the (r = 11 nm) and 120 nm (r = 10 nm, n = 30), re-
c ([001]) and [100] directions. spectively) were observed (Fig. 6a). In contrast,
the presence of an excess of Ca2+ ions ([CO32±]/
2+
calcium alkylbenzenesulfonate surfactants in the bundles [Ca ] = 1:1.6) by incorporation of CaCl2 in the water pools
(EDX analysis and FTIR spectroscopy) suggests that the sepa- reproduced the assembly of the nanofilament bundles
rations between the nanofilaments could correspond to sur- (Fig. 6b).
face-adsorbed layers of surfactant molecules that in turn pre-
vent lateral crystal growth and inter-filament coalescence.
Significantly, electron diffraction patterns recorded from indi-
vidual bundles showed that the vaterite nanofilaments were
single crystals preferentially oriented along the crystallo-
graphic c-axis (Figs. 4b and 4c), indicating that interactions at
the inorganic±surfactant interface were crystallographically
specific under these experimental conditions. In contrast, a
further tenfold increase in the [H2O]/[CaCO3] molar ratio to
3400 gave no nanofilament bundles. Instead, a low yield of irre-
gularly shaped monodisperse vaterite nanoparticles with a
mean size of 130 nm (r = 8 nm, n = 30) was observed by TEM
(Fig. 5). 500nm
Increasing the number ratio of water droplets to ACC nano-
Fig. 5. TEM image of vaterite nanoparticles formed from the transformation of
particles facilitates the transfer of water from the microemul- surfactant-stabilized ACC in NaAOT reverse microemulsions at [H2O]/[Na-
sion droplets, and therefore vaterite nucleation competes more AOT]/[CaCO3] = 3400:340:1.

776 Ó 2002 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim 1616-301X/02/1112-0776 $ 17.50+.50/0 Adv. Funct. Mater. 2002, 12, No. 11±12, December
S. Mann, M. Li/Emergent Nanostructures: Mesoscale Transformation of CaCO3 Nanoparticles

FULL PAPER
a

1000 nm

100nm

Fig. 7. TEM image of mesoporous aragonite produced by the transformation of


surfactant-stabilized ACC nanoparticles in NaAOT microemulsions prepared at
[H2O]/[NaAOT]/[CaCO3] = 680:34:1. Inset shows corresponding powder electron
diffraction with aragonite diffraction rings; d-spacings, 3.38 Š (021), 2.32 Š
(022), 2.11 Š (220), 1.98 Š (221), 1.47 Š (151), 1.41 Š (330) and 1.21 Š (243).

formation occurs. At w = 20, on the other hand, incorporation


of the ACC nanoparticles into water pools with large droplet
diameters (ca. 6.5 nm[10]) results in complete dissolution and
200nm
subsequent bulk crystallization of aragonite. In the presence of
intermediate amounts of bulk waterÐfor example at w = 10
Fig. 6. TEM images of vaterite nanoparticles/nanostructures produced by the
where the droplet diameter is ca. 4.5 nm[10]Ðthe ACC nano-
transformation of surfactant-stabilized ACC nanoparticles in NaAOT micro-
emulsions prepared at [H2O]/[NaAOT]/[CaCO3] = 340:34:1 and in the presence particles become hydrated and coated in a shell of mixed sulfo-
of a molar excess of a) CO32± ([CO32±]/[Ca2+] = 1.6:1), and b) Ca2+ ([CO32±]/ nate surfactants by encapsulation and redistribution within the
[Ca2+] = 1:1.6).
microemulsion water pools. As a consequence, ACC dissolu-
tion and crystal nucleation are not only spatially confined, but
also kinetically controlled by interactions at the surfactant±in-
2.2. Influence of w Value organic interface, which selectively stabilize the vaterite poly-
morph. Moreover, as described above, coupling between the
The amount of water trapped within individual droplets had kinetics of phase transformation and the subsequent growth
a marked effect both on the polymorph structure and architec- and aggregation of the surfactant-coated vaterite nanoparticles
ture of the calcium carbonate materials produced by micro- results in higher-order structures, such as spherical, spindle-
emulsion-mediated phase transformation of the surfactant-sta- shaped or filamentous aggregates.
bilized ACC nanoparticles. In contrast to the vaterite
nanostructures described above for w = 10, ACC transforma-
tion was arrested in the absence of bulk water (reverse mi- 3. Conclusions and General Model
celles, w < 3), such that the TEM images showed only discrete
electron-dense 5 nm sized amorphous cores. On the other The above results indicate that a range of hybrid surfactant±
hand, experiments undertaken at w = 20 and [H2O]/ vaterite nanostructures can be produced through microemul-
[CaCO3] = 680:1 gave a mesoporous 3D sponge-like architec- sion-mediated phase transformation of surfactant-stabilized
ture of interconnected crystalline nanofilaments, ca. 10 nm in ACC nanoparticles, by controlling the number of water drop-
width (Fig. 7). Significantly, powder X-ray diffraction and elec- lets per ACC nanoparticle at a constant w value of 10 (Fig. 8).
tron diffraction studies showed that the inorganic network ob- Interactions between the incipient vaterite nuclei and the sur-
tained after 2 days consisted of a polycrystalline array of arago- rounding alkylbenzenesulfonate and AOT anionic surfactants
nite crystals (Fig. 7, inset). are likely to be mediated by the surface charge density, which
The results suggest that the nanostructures originate from a in turn is related to the surface structure, degree of hydration
dissolution±reprecipitation process that is locally confined and and ionic strength of the boundary layers. In this regard, the
dependent on the relative solubilities and surface stabilities of spindle-shaped aggregates shown in Figure 3 represent an
the ACC precursor and crystalline product. The low water con- intermediate stage between the hydrophobic spheroidal clus-
tent of the reverse micelles is presumably not sufficient to dis- ters and highly anisotropic nanofilament bundles. Some limited
solve the surfactant-stabilized ACC phase, and no phase trans- growth in the vaterite nanoparticles is achieved due to the

Adv. Funct. Mater. 2002, 12, No. 11±12, December Ó 2002 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim 1616-301X/02/1112-0777 $ 17.50+.50/0 777
S. Mann, M. Li/Emergent Nanostructures: Mesoscale Transformation of CaCO3 Nanoparticles
FULL PAPER

w = 10 ping of the surface, face-specific


growth, and unrestricted bulk growth
H2O/CaCO3 associated with facile water-droplet
exchange and expansion, which de-
+ H2 O pends on the nature and strength of
34 the interactions at the surfactant±inor-
ganic interface. Moreover, these inter-
actions can operate synergistically to
produce organized hybrid nanostruc-
alkylbenzenesulfonate AOT
tures if the shape anisotropy exhibited
170
amorphous CaCO3 by the habit-modified nanoparticles is
coupled in situ with long-range self-as-
crystalline CaCO3
sembly of the surfactant molecules.
Although only a narrow window of
340 conditions is likely to be concomitant
H2O/AOT/CaCO3
+ CO3 2- in general with the emergence of
= 680 : 34 : 1
structural and morphological complex-
ity, understanding the factors that gov-
ern these special cases has important
3400 generic consequences for the repro-
200 nm ducible synthesis of nanostructured
hybrid materials with vectorially con-
trolled architectures.
aragonite sponge vaterite-surfactant nanostructures
Finally, we note that the above
Fig. 8. General scheme showing experimental conditions and types of hybrid surfactant-vaterite nanostructures results could offer new insights for
synthesized by microemulsion-mediated phase transformation of surfactant-stabilized ACC nanoparticles. Reac-
tants are shown top/middle left. Nanostructures produced at w = 10 using different water droplet/amorphous biomineralization mechanisms. Sur-
CaCO3 nanoparticle ratios are shown on the right. The aggregated structures are drawn approximately to scale prisingly, ACC is present as a stable
(bar = 200 nm).
biomineral in the body and tunic spic-
ules of ascidians,[26,27] in skeletal spic-
increased level of hydration in the system, but the nanoparti- ules of calcareous sponges,[26] plant cystoliths,[28] crustacean
cles are significantly less elongated than in the nanofilament exoskeletons,[29,30] and shells of certain freshwater snails.[31]
bundles, probably because the surfactants remain strongly an- The amorphous mineral is associated with polysaccharides and
chored to all the crystal surfaces. This is reduced for the nanofi- proteins specifically enriched in glutamic acid, serine, and
laments because of the increased hydration, but importantly, threonine,[26,27] and/or inorganic ions such as Mg2+ and phos-
only to a level where the surfactant molecules appear to be phate,[27] and model studies have confirmed that these addi-
preferentially bound to crystal faces parallel to the vaterite tives can stabilize the formation of ACC under laboratory con-
c-axis. Further increases in the water content removes this face ditions.[27,32] In contrast, the deposition of calcite in sea urchin
selectivity to give large non-modified vaterite crystals as shown larval spicules[33,34] and aragonite in a freshwater snail shell[31]
in Figure 5. occurs by controlled transformation of a transient ACC inter-
Besides the degree of hydration, interfacial interactions be- mediate, and a similar role for ACC is proposed to explain the
tween the vaterite nanocrystallites and sulfonate surfactants unusually high magnesium levels associated with a wide range
can be fine-tuned by modifications in the [CO32±]/[Ca2+] molar of calcitic biominerals.[32] Because biominerals are produced
ratio (see Fig. 6). This is consistent with previous work on from solutions that contain a heterogeneous mixture of organic
BaSO4 and BaCrO4 crystallization in AOT microemul- molecules and ions, kinetic factors usually determine the onset
sions,[10,14] which showed that nanofilament bundles, linear of primary nucleation,[11] and ACC precursors in association
chains, or discrete spherical nanoparticles could be synthesized with surface-adsorbed organic molecules are likely to be the
depending on whether the anion/cation molar ratio was < 1, rule rather than exception in calcium carbonate biomineraliza-
1 or > 1, respectively. Analogous with the previously proposed tion. Moreover, as discussed above, depending on the nature
mechanism,[10] an excess of CO32± induces a net negative sur- and strength of interactions at the inorganic±organic interface,
face charge at the interface with the growing nanoparticles, transformation of the amorphous primary particles can become
such that electrostatic interactions with the anionic headgroups vectorially regulated to produce organized hybrid nanostruc-
of AOT and alkylbenzenesulfonate are significantly reduced, tures that are crystallographically oriented. Thus it seems pos-
and the vaterite crystals develop relatively unperturbed. On sible that the oriented nucleation and growth of biominerals
the other hand, molar equivalence or an excess of Ca2+ ions in could be determined by co-operative processes involving phase
the water pools results in positively charged inorganic surfaces transformations at the mesoscopic level, as well as by molecu-
that interact strongly with the surfactant headgroups. Thus, lar-based mechanisms such as described in classical theories of
there appears to be a delicate balance between complete cap- organic matrix-mediated epitaxy.[35,36]

778 Ó 2002 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim 1616-301X/02/1112-0778 $ 17.50+.50/0 Adv. Funct. Mater. 2002, 12, No. 11±12, December
S. Mann, M. Li/Emergent Nanostructures: Mesoscale Transformation of CaCO3 Nanoparticles

FULL PAPER
4. Experimental ±
[1] P. Yang, A. H. Rizvi, B. Messer, B. F. Chmelka, G. M. Whitesides, G. D.
Stucky, Adv. Mater. 2001, 13, 427.
Materials: Analytical-grade sodium bis(2-ethylhexyl)sulfosuccinate (NaAOT) [2] W. Shenton, D. Pum, U. B. Sleytr, S. Mann, Nature 1998, 389, 585.
and isooctane (2,2,4-trimethylpentane) were purchased from BDH and Aldrich, [3] C. B. Murray, C. R. Kagan, M. G. Bawendi, Science 1995, 270, 1335.
respectively, and used without further purification. Surfactant-stabilized ACC [4] Z. L. Wang, Adv. Mater. 1998, 10, 13.
nanoparticles were a gift from Dr. J. Eastoe (University of Bristol). The sample, [5] R. P. Andres, J. D. Bielefeld, J. I. Henderson, D. B. Janes, V. R. Kolagunta,
which was dried by rotary evaporation, was similar to the ¢V-series¢ characterized C. P. Kubiak, W. J. Mahoney, R. G. Osifchin, Science 1996, 273, 1690.
previously [21]. CHN elemental analysis (total C = 41.75 %, N = 0 %, H = 5.89 %, [6] M. Brust, D. Bethell, D. J. Schiffrin, C. J. Kiely, Adv. Mater. 1995, 7, 795.
inorganic C = 3.46 %) and thermogravimetric analysis (Simultaneous Thermal [7] C. A. Mirkin, R. L. Letsinger, R. C. Mucic, J. J. Storhoff, Nature 1996, 382,
Analysis STA409EP) indicated that the nanoparticles contained 29 % (w/w) 607.
CaCO3. Fourier transform infrared (FTIR) spectroscopy showed bands at [8] S. Mann, W. Shenton, M. Li, S. Connolly, D. Fitzmaurice, Adv. Mater. 2000,
863 cm±1 and 1079 cm±1 corresponding to the CO32± anion in amorphous CaCO3, 12, 147.
as well as at 1495 cm±1 (C=C aromatic), 2958, 2927, and 2856 cm±1 (C±H [9] E. Dujardin, L.-B. Hsin, C. R. C. Wang, S. Mann, Chem. Commun. 2001,
stretches), and 1183 cm±1 (O=S=O symmetric stretch). Suspensions of alkylben- 1264.
zenesulfonate-capped ACC nanoparticles in pure toluene were transparent [10] M. Li, H. Schnablegger, S. Mann, Nature 1999, 402, 393.
stable solutions, which when air-dried onto TEM grids showed non-aggregated [11] S. Mann, Biomineralization. Principles and Concepts in Bioinorganic Mate-
electron dense amorphous cores, ca. 5 nm in diameter. rials Chemistry, Oxford University Press, Oxford, UK 2001.
Methods: Water-in-oil NaAOT microemulsions were prepared by adding small [12] H. A. Lowenstam, S. Weiner, On Biomineralization, Oxford University
amounts of distilled, de-ionised water (pH 7.3) to NaAOT isooctane solutions Press, New York. 1989.
(0.1 M in isooctane) to give w values (w = [H2O]/[NaAOT]) of between 1 and 20. [13] S. Mann, Angew. Chem. Int. Ed. 2000, 39, 3392.
Typically, 0.18 mL of water was added to 10 mL of the 0.1 M NaAOT solution to [14] M. Li, S. Mann, Langmuir 2000, 16, 7088.
give a w value of 10, which was used for most of the experiments. Different [15] L. Qi, J. Ma, H. Cheng, Z. Zhao, J. Phys. Chem. B 1997, 101, 3460.
amounts of dried samples of the alkylbenzenesulfonate-capped ACC nanoparti- [16] G. D. Rees, R. Evans-Gowing, S. J. Hammond, B. H. Robinson, Langmuir,
cles were added to a constant microemulsion volume to give CaCO3 final concen- 1999, 15, 1993.
trations between 0.3 and 30 mM, and corresponding [H2O]/[CaCO3] molar ratios [17] M. Antonietti, M. Breulmann, C. G. Göltner, H. Cölfen, K. K. W. Wong,
of 34, 170, 340, 680, and 3400, each at w = 10. For example, addition of 100 mg of D. Walsh, S. Mann, Chem. Eur. J. 1998, 4, 2491.
ACC to 10 mL of a NaAOT microemulsion (0.1 M in isooctane, w = 10) gave a [18] R. Kniep, S. Busch, Angew. Chem. Int. Ed. 1996, 35, 2624.
CaCO3 concentration of 30 mM (measured against total microemulsion volume) [19] J. M. Garcia-Ruiz, J. Cryst. Growth 1985, 73, 251.
and [H2O]/[NaAOT]/[CaCO3] = 34:3.4:1. The dispersions were shaken vigorously [20] L. Qi, H. Cölfen, M. Antonietti, M. Li, J. D. Hopwood, A. J. Ashley,
and then left to stand at room temperature, 4 C or 50 C without stirring. Sam- S. Mann, Chem. Eur. J. 2001, 7, 3526.
ples for analysis were taken after periods between 10 min and two months. [21] I. Markovic, R. H. Ottewill, Colloid Polym. Sci. 1986, 264, 454.
Similar experiments were undertaken using NaAOT microemulsions (w = 10) [22] D. C. Steyler, B. H. Robinson, J. Eastoe, K. Ibel, J. C. Dore, I. MacDonald,
prepared from; a) water droplets at different pH values (between 3 and 11), Langmuir. 1993, 9, 903.
b) droplets containing aqueous Na2CO3 (100 mM, pH 11.75; molar ratios, [H2O]/ [23] L. Brevevic, A. Nielson, J. Cryst. Growth 1989, 98, 504.
[NaAOT]/[CaCO3] = 340:34:1, [CO32±]/[Ca2+] = 1.6:1, and c) droplets containing [24] M. M. Reddy, G. H. Nancollas, J. Cryst. Growth 1976, 35, 33.
aqueous CaCl2 (100 mM, pH 6.7; molar ratios, [H2O]/[NaAOT]/ [25] S. L. Tracey, D. A. Williams, H. M. Jennings, J. Cryst. Growth 1998, 193,
[CaCO3] = 340:34:1, [CO32±]/[Ca2+] = 1:1.6). Control experiments were carried out 382.
by adding dried samples of the surfactant-stabilized ACC to NaAOT microemul- [26] J. Aizenberg, G. Lambert, L. Addadi, S. Weiner, Adv. Mater. 1996, 8, 222.
sions in the absence of added water (reverse micelles). [27] J. Aizenberg, G. Lambert, S. Weiner, L. Addadi, J. Am. Chem. Soc. 2002,
Samples for transmission electron microscopy (TEM), electron diffraction and 124, 32.
EDX analysis were deposited onto carbon-coated, 3 mm diameter, copper [28] M. G. Taylor, K. Simkiss, G. N. Greaves, M. Okazaki, S. Mann, Proc. R. Soc.
electron microscope grids. After drying in air, the grids were washed with pure London, Ser. B 1993, 252, 75.
isooctane. TEM analysis was performed in bright-field mode using a JEOL 1200 [29] A. P. Vinogradov, The Elementary Chemical Composition of Marine Organ-
EX electron microscope operating at 120 keV and linked with an Oxford Instru- isms, Sears Foundation for Marine Research, New Haven, CT 1953.
ments X-ray analysis system. Samples for scanning electron microscopy (SEM) [30] Y. Levi-Kalisman, S. Raz, S. Weiner, L. Addadi, Adv. Funct. Mater. 2002,
were mounted onto circular aluminum stubs or carbon adhesive pads attached 12, 43.
to aluminum stubs. Samples were air-dried and gold-coated with an Edwards [31] B. Hasse, H. Ehrenberg, J. C. Marxen, W. Becker, M. Epple, Chem. Eur.
S150B sputter-coater or left uncoated. SEM analyses were carried out using J. 2000, 6, 3679.
JEOL JSM 5600LV and JEOL JSM 6300F FEGSEM scanning electron micro- [32] S. Raz, S. Weiner, L. Addadi, Adv. Mater. 2000, 12, 38.
scopes operating at accelerating voltages of 1±30 keV and 0.5±30 keV, respective- [33] E. Beniash, J. Aizenberg, L. Addadi, S. Weiner, Proc. R. Soc. London,
ly. Samples were studied by Fourier transform infrared (FTIR) spectroscopy Ser. B 1997, 264, 461.
using KBr disks. [34] E. Beniash, L. Addadi, S. Weiner, J. Struct. Biol. 1999, 125, 50.
[35] S. Mann, Nature 1988, 332, 119.
Received: April 4, 2002 [36] L. Addadi, J. Moradian, E Shay, N. G. Maroudas, S. Weiner, Proc. Natl.
Final version: June 25, 2002 Acad. Sci. USA 1987, 84, 2732.

______________________

Adv. Funct. Mater. 2002, 12, No. 11±12, December Ó 2002 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim 1616-301X/02/1112-0779 $ 17.50+.50/0 779

Potrebbero piacerti anche