SlideShare a Scribd company logo
1 of 51
Download to read offline
© 2010 WILEY-VCH Verlag GmbH & Co. KGaA, WeinheimAdv. Mater. 2010, 22, 3729–3734 3729
www.advmat.de
www.MaterialsViews.com
COMMUNICATION
By Halei Zhai, Wenge Jiang, Jinhui Tao, Siyi Lin, Xiaobin Chu, Xurong Xu,
and Ruikang Tang*
Self-Assembled Organic–Inorganic Hybrid Elastic Crystal
via Biomimetic Mineralization
[*] H. Zhai, W. Jiang, J. Tao, S. Lin, X. Chu, Dr. X. Xu, Prof. R. Tang
Center for Biomaterials and Biopathways
and Department of Chemistry
Zhejiang University
Hangzhou, 310027 (P.R. China)
E-mail: rtang@zju.edu.cn
Dr. X. Xu, Prof. R. Tang
State Key Laboratory of Silicon Materials
Zhejiang University
Hangzhou, 310027 (P.R. China)
DOI: 10.1002/adma.201000941
It is generally accepted that biomaterials have unique physi-
cochemical properties.[1] Inspired by biological systems, sci-
entists have been studying biomimetic methods to fabricate
functional materials.[2] Almost all biomaterials possess a
common multi-component feature.[1,3] These composites fre-
quently have ordered organic–inorganic hybrid structures and
their properties are distinct from the individual components.
For example, in a multilayered complex of inorganic aragonite
tablets and organic substrate, the fracture toughness of nacre is
significantly improved to three thousand times greater than that
of synthetic aragonite.[4] Another striking example is biological
bone. In bone, the hydroxyapatite (HAP) phase crystallizes in
the nanoscaled channels formed by the staggered alignment
of the protein matrix. The typical HAP crystals in bone are
plate-shaped with extremely thin thickness (1.5–2 nm), which
is the smallest known dimension of the biologically formed
crystals.[5] In nature the organic and inorganic components inti-
mately associate into well-organized hybrid structures to ensure
optimal strength and flexural stress.[6,7] Therefore, in biomi-
metic designs and fabrications the formation of such ordered
nanostructures is a key challenge.
The formation of inorganic crystals in living organisms is
regulated by the organic matrix. Generally, different organic
templates and additives lead to variety in the morphology, size,
orientation, and assembly of the inorganic crystal by medi-
ating its nucleation and growth.[8,9] Although many organic–
inorganic nanocomposites have been reported,[10] the self-
formation of ultrathin organic–inorganic substructures is still
difficult to achieve by using a simple bottom-up approach. But
the self-formed ordered and intimate combination of organic
additives and inorganic crystals at the nanoscale is a crucial
requirement for bioactive composites.[11] Here we prepare an
organic–inorganic hybrid crystal by the self-assembly of cal-
cium phosphate, surfactant, and protein. This hybrid crystal is
composed of uniform and alternate organic–inorganic layers
at the nanoscale. Both the inorganic crystalline phase and
organic phases in the hybrid crystals have an ultrathin thick-
ness of 1–2 nm. The two ordered components form simultane-
ously during the crystal generation so that they integrate well
with each other to form a superstructure. It is of great impor-
tance that such biomimetic crystals are considerably flexible
and elastic.
It is believed that functional organic molecules can interact
with calcium species at the organic–inorganic interfaces to
modulate the growth and assembly process of the inorganic
crystals. The globular protein bovine serum albumin (BSA),
which comprises a single chain of 583 amino acid residues,
is one of the most studied proteins. It is widely used as a
model protein in many fields including biomimetic miner-
alization.[12] Surfactants are widely applied as the crystalliza-
tion templates in many biomimetic studies.[13] However, the
cooperation of different organic additives has been frequently
overlooked in previous works because of the complicated inter-
actions in the system.[14] Actually, the interactions of a sur-
factant molecule and protein are widely found in biological
systems, for example, the interaction of protein with cell mem-
brane surfactants. The selected two compounds can represent
the protein matrix and special small functional molecules in
biomimetic mineralization studies. Usually, proteins and sur-
factants can form complexes in solution, which are frequently
described by a “necklace bead model”. The micelle-like clus-
ters of surfactants scatter along the polypeptide chains like the
pearls in a necklace.[15] The hydrophilic groups of micelles
are exposed to aqueous solutions and their configuration can
be adjusted. In such protein–surfactant complexes, the protein
is functionalized by the surfactant; meanwhile the aggrega-
tion behavior of the surfactant is also affected by the protein
structure. Here we find that the complex of BSA and an ani-
onic surfactant (sodium bis(2-ethylhexyl) sulfosuccinate, AOT)
could self-assemble into regular rhombus plates with a spe-
cific organic–inorganic substructure in a calcium phosphate
solution.
Scanning electron microscopy (SEM) shows the uniform
rhombic plates formed by the collaboration of calcium phos-
phate, BSA, and AOT (Figure 1a).The typical rhombs are
300–400 nm in the long axis and 200–300 nm in the short axis.
Their typical thickness is 80–100 nm. These rhombs are stable
and their structures can endure in solution or in air for months.
The energy-dispersive X-ray spectroscopy (EDS) reveals the pres-
ence of calcium and phosphate ions in the rhombs; the atomic
ratio of Ca:P is around 1.5. In addition to the elements of C
and O, S was also detected (Figure S1 and Table S1 of the Sup-
porting Information), indicating the presence of AOT (–SO3
2−).
The organic–inorganic hybrid composite was also confirmed by
3730
www.advmat.de
www.MaterialsViews.com
© 2010 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim Adv. Mater. 2010, 22, 3729–3734
COMMUNICATION
characteristic substructure: two independent sets of diffrac-
tion peaks were detected by using wide-angle X-ray diffraction
(WAXD) and small-angle X-ray diffraction (SAXD) (Figure 1e).
In the small-angle region, a typical reflection characteristic
of lamellar structures is observed. The interspacing distance,
d = 3.12 nm, was calculated by using the reflection peak at
2θ = 2.83° ((001) reflection of the rhomb crystals). The (002)
and (003) reflections were detected at 2θ = 5.71° (d = 1.55 nm)
in SAXD and 2θ = 8.45° (d = 1.05 nm) in WAXD, respectively.
These sharp peaks show the rhombs had a highly ordered
lamellar structure. The other WAXD peaks in the normal range
(2θ > 10°) indicate that the crystallized mineral phase is a HAP-
like phase. These examinations clearly demonstrate that there
are two independent lattice structures within a rhomb crystal. It
is important that the organic and inorganic phases are orderly
arranged to form the hybrid materials rather than the simple
and disordered mixture. By using a side view of the ultrathin
Fourier transform infrared spectroscopy (FTIR). The peaks at
1737, 1459, and 1419 cm−1 are the characteristic peaks of AOT,
while the bands at 1655 (amide I) and 1553 cm−1 (amide II) indi-
cate the presence of BSA. In addition, the broad peaks at 1023 and
567cm−1 areduetothepresenceoftheinorganicphosphategroup
(Figure 1b). Thermogravimetric analysis (TGA) and differential
scanning calorimetry (DSC) showed the presence of 21.4 %
organic component (the organics decomposed at temperatures
of 200–500 °C) and 62.1 % inorganic composite (the residue at
temperatures above 500 °C, Figure 1c). From these results, we
can conclude that the rhombs are the hybrid materials of inor-
ganic (calcium phosphate) and organic phases (BSA and AOT).
The regular rhombs were examined by means of trans-
mission electron microscopy (TEM, Figure 1d). The selected
area electron diffraction (SAED) pattern shows the inor-
ganic phase is in a crystalline form and the pattern is similar
to that of HAP tiny crystallites. Abnormally, the crystal has a
Figure 1. a) SEM image of the rhombs. Inset: enlargement of the rhomb in the white circle. b) FTIR curves of the rhombs (bottom) and AOT (top).
The characteristic peaks for BSA, AOT, and phosphate, are marked as circles, triangles, squares, respectively. c) TGA and DSC analysis under a nitrogen
atmosphere. The weight percentages of water and organic component are labeled. d) Transmission electron microscopy (TEM) image of the rhombs.
Inset: selected area electron diffraction (SAED) pattern corresponding to the white circled area. e) Wide-angle and small-angle (inset) X-ray diffraction
(WAXD and SAXD, respectively) patterns of the rhombs. f) TEM side view of an ultrathin sectioned rhomb.
3731
www.advmat.de
www.MaterialsViews.com
© 2010 WILEY-VCH Verlag GmbH & Co. KGaA, WeinheimAdv. Mater. 2010, 22, 3729–3734
COMMUNICATION
a strong binding effect with calcium ions as
a result of the highly charged –SO3
2−
groups
(Figure S2, Supporting Information). But the
interaction between calcium and BSA was rel-
atively poor. Since AOT molecules aggregated
onto the BSA chains according to the “neck-
lace bead model”, the local concentrations of
calcium around the BSA–AOT complex were
greater than that in the bulk solution so that
the AOT aggregates on BSA provided the het-
erogeneous nucleation sites for calcium phos-
phate. Moreover, the AOT molecules were
organized by the BSA structure so that the
complexes could induce the ordered assembly
of calcium phosphate. We suggest that the
mineral surfaces also act as the stable solid
substrates for the self-assembly of the BSA–
AOT complex. Thus, the lamellar organic–
inorganic structures could be bottom-up
assembled in the solutions spontaneously.
Accordingly, the substructure of the hybrid
rhombs is the alternate combination of the
ultrathin nanocrystal layer and the BSA–
AOT monolayer (Figure 2b), which is analo-
gous to the nanoscale characteristics of many
natural hybrid composites.[1,3,6,7]
Structured materials are usually asso-
ciated with unique physicochemical and
biological properties.[16]
Both advantages
of inorganic and organic phases can be
present in one hybrid material if these two
components can be well-integrated at the
nanoscale.[17]
Although the main compo-
nent of the rhombs is the crystallized cal-
cium phosphate, a rigid inorganic phase,
flexile and elastic behavior of the hybrid
crystal was obtained. Figure 3a illustrates a side view of a
rhomb: the whole crystal and its organic–inorganic layers are
bent to some extent. Interestingly, a similar bent wave shape
can be seen in the typical organic–inorganic hybrid reinforced
materials such as some polymer–clay nanocoposites.[18]
In
order to confirm the mechanical features of the material, a
force curve examination using atomic force microscopy (AFM,
Figure 3b) was applied. The cantilever was very sensitive to
the tip force and its deflection curve could qualitatively repre-
sent the hardness of the examined surface. In contrast to the
typical sudden and straight force–deflection lines for the rigid
silicon substrate (modulus of 130 GPa, which is similar to that
of pure HAP crystals: 112 GPa[19]
), the loading force increased
smoothly with an increase of the deflection degree of the
AFM cantilever. The buffer effect in the AFM force examina-
tion indicates that the rhombs are not rigid. This characteristic
was similar to that of a typical soft material, polystyrene (PS,
modulus of about 3 GPa). It is interesting that no obvious per-
manent damage or indention point was detected on the rhomb
surface after the loading–unloading cycles (inset of Figure 3b)
in the AFM examination. In order to quantitatively understand
the mechanical properties of the hybrid, a nanoindentation
measurement with a diamond indenter tip was additonally
section of the rhombs under TEM, the lamellar structure is
shown in Figure 1f: the dark region corresponds to the inor-
ganic phase (crystallized calcium phosphate) and the light one is
the organic phase. The individual organic and inorganic phases
are alternately stacked. Each layer structure could be identified
readily at the nanoscale in the hybrid crystal. These two distinct
units are well integrated so that the complete hybrid crystals
can be finally produced at the nanoscale. The thickness of each
organic–inorganic unit is about 3.2 ± 0.2 nm, which is in good
agreement with the calculated d value from the SAXD study.
It is noted that the thickenss of the mineral layer is only about
2 nm; this dimension is close to that of biological ultrathin
HAP crystallites formed between the collagen fibers of bone.
In order to understand the substructure of the rhombs, the
organic component was partially degraded by a 5 % NaOCl solu-
tion. Thus, the mineral layer in the complex could be observed
directly by TEM (Figure 2a). Small crystalline platelets, tens of
nanometers in dimension (length and width), were frequently
observed. In a rhomb crystal, the locations of inorganic crystal-
line platelets are restricted by the adjacent protein–BSA organic
frames. Thus, the continuous inorganic ultrathin layers might
be formed between the frames by using the nanocrystallites.
The conductivity investigations showed that AOT molecules had
Figure 2. a) TEM image of the rhombs etched by 5 % NaOCl; The inset is its fast Fourier trans-
form (FFT) image. b) Substructures of the organic–inorganic rhombs. AOT: small molecules
with round head; BSA: long dark chains; mineral phase: rectangles.
3732
www.advmat.de
www.MaterialsViews.com
© 2010 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim Adv. Mater. 2010, 22, 3729–3734
COMMUNICATION
could even partially recover during the unloading processes.
In contrast, the unloading curves should be vertical if the solid
phase was rigid.[20] Since the indentation depth was greater
than 20 % of the sample thickness, the Bec model[21]
for a thin
soft material on a hard substrate was applied in the estimation
of the modulus (see details in Supporting Information). By
using the loading–unloading curves, the calculated modulus of
the organic–inorganic rhombs was 6.64 ± 1.41 GPa. This value
was even lower than the modulus of elastic-featured human
vertebral trabeculae, 13.5 ± 2.0 GPa.[22] Similar to biological
bone, both the elastic and hardness features were successfully
integrated by the nanostructured assembly of organic and inor-
ganic ultrathin phases, implying that the hybrid rhombs resolve
the brittleness shortage of inorganic crystals and improve the
material’s toughness. Actually, this is a smart and important
strategy of living organisms to generate functional biomaterials
by means of hybrid nanostructures.
Many research efforts often focus on the controlling effect
of the organic matrix on inorganic mineralization processes,
which mediates the size, morphology, and orientation of inor-
ganic crystals. Such an understanding implies a one-way con-
trol of inorganic phase formation by organic additives. Thus,
the organic templates are often required prior to the controlled
crystallization in order to obtain hybrid materials. However,
this understanding is not suitable in the current case. It was
noted that the BSA–AOT complexes could not form the rhomb
structure spontaneously in calcium solutions. Neither our
experiment nor the published literature detected the BSA–AOT
rhomb in the absence of any mineral phase. Only poorly crystal-
line calcium phosphate spherical particles were obtained if only
BSA was added into the calcium phosphate solution. Besides,
AOT alone resulted in the conventional rod-like HAP crystals
(Figure S3, Supporting Information) without any substructure.
Clearly, the formation of the hybrid rhombs is attributed to the
coexistence of BSA, AOT, and calcium phosphate, which is an
emergent process. As mentioned above, the presence of the
inorganic part also induces the assembly and structure of the
organic components during mineralization.[23] Additionally,
the changes of BSA and AOT concentrations within a certain
range only affects the size and morphology of the resultant
rhombs (Figure S4, Supporting Information). However, their
internal substructure was not altered at all (Figure S5). This
phenomenon could be explained by the regulation effect of
surfactant on the complex assembly, which has been demon-
strated by previous work.[15]
We noted that the assembly process rather than conven-
tional crystal growth occurred in the rhomb formation. No
obvious signal between 50 and 100 nm was observed during
the whole reaction process by dynamic light scattering (DLS,
Figure 4). At the initial stage of crystallization, two individual
distribution peaks existed in the DLS pattern. The small one
(∼20 nm) represented the BSA–AOT building block in the reac-
tion solution (Figure S6, Supporting Information), while the
large one (∼300 nm) belonged to the final product. The frac-
tion of the building block decreased gradually with the reaction,
while the intensity of the product increased. Eventually, only
the final product could be found at the end of the experiment.
The product size did not increase during the reaction. Accord-
ingly, the ex situ electron microscopy studies also demonstrated
performed on the rhombs so that the modulus of the material
could be calculated.[20] The solid and dashed lines represent the
loading and unloading processes, respectively (Figure 3c). The
relatively great indentation depth with different loading forces
from 25 to 40 μN were used to demonstrate the elastic charac-
teristic of the whole nanoplate well. Under such great external
forces, the deformation of the plates was significant. However,
the thin crystals did not collapsed and the depressed surfaces
Figure 3. a) TEM image of ultrathin section of the rhomb. b) Atomic force
microscopy (AFM) force curves of silicon substrate, rhombs (Rh) and poly-
styrene (PS). Cantilever deflection represents the deformation distance
of the sensitive AFM cantilever. Inset: AFM image of the plate after the
loading–unloading cycle. c) The nanoindentation curves of rhombs. The
displacement here means the indentation distance from the surface.
3733
www.advmat.de
www.MaterialsViews.com
© 2010 WILEY-VCH Verlag GmbH & Co. KGaA, WeinheimAdv. Mater. 2010, 22, 3729–3734
COMMUNICATION
and nanoindentation were prepared by spin-coating 100 μL of slurry on
silicon wafers (3000 rpm). For ultrathin-sectioned TEM examination,
rhombs were embedded in 0.5 mL of epoxy. The mixture was solidified
at 80 °C for 12 h and then carefully microtomed by a Reichert–Jung
Ultracut E using a diamond knife. The typical thickness of the ultrathin
sections was ∼80 nm.
Characterization: SEM was performed by using a HITACHI S-4800
microscope at an accelerating voltage of 5 kV. FTIR spectroscopy (Nicolet
Nexus 670) was used to determine the composition of the products.
Thermogravimetric analysis was carried out by a TA Instrument SDT
Q600. The experiment was measured in a temperature range of 22–600 °C
at a heating rate of 10 °C min−1 under nitrogen atmosphere. TEM
observations were performed by a Philips CM200UT microscope at a
typical accelerating voltage of 160 kV. WAXD and SAXD were carried out
by means of a Rigaku D/max-2550pc instrument with monochromatized
Cu Kα radiation and a scanning step of 0.02°. AFM images were
collected by a Veeco multimode scanning probe microscope with Nano
IVa controller. The measurements were performed using an E head and
a silica tip (Veeco) on a cantilever with a spring constant of 40 N m−1
in tapping mode with filters off, with a scanning rate of 20−60 Hz. The
qualitative measurement of the mechanical properties was performed
by the cantilever deflection in the AFM force curve. The data was
collected for 200 individual force curves on 10 different rhombs. The
nanoindentation measurements were performed by a Tribo-Indenter
In-Situ Nanomechanical Test System with a Berkovich diamond indenter
(tip radius of about 50 nm). The system was calibrated by using fused
quartz before indentation. The data was collected using TestWorks 4
(MTS Systems). The modulus was calculated using the Oliver and Pharr
method and the substrate effect was corrected by the Bec model. The
DLS measurements were taken by using a Brookhaven Instruments 90
Plus particle size analyzer. Conductivity measurements were carried out
by Conducometer DDS-11A at 30 °C. The conductivity electrode was
calibrated using 0.01 M KCl solution prior to use.
Supporting Information
Supporting Information is available online from Wiley InterScience or
from the author.
Acknowledgements
We thank Haihua Pan and Yuan Su for their helpful discussions, Yuewen
Wang, Jieru Wang, Yin Xu, and Xiaoming Tang for assistance in material
characterization techniques. This work was supported by the National
Natural Science Foundation of China (20601023 and 20871102),
Zhejiang Provincial Natural Science Foundation (R407087), the
Fundamental Research Funds for the Central Universities and Daming
Biomineralization Foundation.
Received: March 16, 2010
Revised: April 5, 2010
Published online: July 21, 2010
the absence of intermediate solid or phase during the growth.
The DLS result reveals an abnormal pathway in the organic–
inorganic hybrid material assembly. We suppose that the BSA–
AOT complexes induce the mineral crystallization firstly and
then they are restructured by the mineral phase to form the
alternative layer structure by a cooperative effect. However, the
detailed mechanism needs further investigation.
In this Communication we demonstrate that organic–inor-
ganic hybrid rhombs with a lamellar superstructure can be
self-generated by protein, surfactant molecules, and mineral
phases. Each crystal contains two basic nanoscaled subunits: the
ultrathin inorganic mineral and organic ultrathin layers. These
layers are formed simultaneously and integrate well by self-
assembly to generate the hybrid crystals. During this process
the cooperative effect between the organic and inorganic phases
is key. The ordered organic–inorganic nanostructure confers the
optimum mechanical properties on the resultant hybrid mate-
rial. The current study provides further evidence of the biomi-
metic fabrication of functional materials.
Experimental Section
Materials: Triply distilled CO2-free water was used in the experiment.
Ca(NO3)2 and (NH4)2HPO4 were of analytical grade and their solutions
were filtered twice through Millipore films (0.22 μm) prior to use. BSA
(Albumin Bovine fraction V, BR, purity >98 %) and AOT (Aldrich) were
used directly without further purification.
Preparation: An aqueous solution (100 mL) containing AOT (4 mM)
and BSA (1 mg mL−1
) was prepared. The solution pH was adjusted to
10.0 ± 0.5 at room temperature by ammonia solution (3 M). Ca(NO3)2
solution (50 mL, 5mM, pH = 10.0 ± 0.5) was added to the mixed solution
at a rate of 10 mL min−1
and the solution was stired for 30 min. After that,
(NH4)2HPO4 solution (50 mL, 5mM, pH = 10.0 ± 0.5) was added dropwise
at a rate of 1.5 mL min−1. The slurry was examined by DLS periodically
and the formed solids were collected by high-speed centrifugation at
10 000 rpm. All the solid samples were washed by water three times
and were vacuum-dried at 35 ± 1 °C. Freshly prepared rhombs were
dispersed in ethanol (∼0.5 mg mL−1
) and collected on carbon-coated
copper grids for TEM examination. Samples for AFM measurements
Figure 4. Dynamic light scattering (DLS) size distribution curves at dif-
ferent stages during the emergent formation of rhombs. The percentage
values are calculated by using the statistics of the particle amounts.
[1] S. Mann, Biomineralization: Principles and Concepts in Bioinorganic
Materials Chemistry, Oxford University Press, Oxford 2001.
[2] a) L. P. Lee, R. Szema, Science 2005, 310, 1148; b) C. Sanchez, H.
Arribart, M. M. G. Guille, Nat. Mater. 2005, 4, 277; c) T. Kato, A.
Sugawara, N. Hosoda, Adv. Mater. 2002, 14, 869; d) T. Sun, L. Feng,
X. Gao, L. Jiang, Acc. Chem. Res. 2005, 38, 644.
[3] a) N. Watabe, J. Ultrastruct. Res. 1965, 12, 351; b) L. C. Palmer, C. J.
Newcomb, S. R. Kaltz, E. D. Spoerke, S. I. Stupp, Chem. Rev. 2008,
108, 4754; c) H. O. Fabritius, C. Sachs, P. R. Triguero, D. Roobe,
Adv. Mater. 2009, 21, 391.
3734
www.advmat.de
www.MaterialsViews.com
© 2010 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim Adv. Mater. 2010, 22, 3729–3734
COMMUNICATION
2005, 285, 562; d) S. Chodankar, V. K. Aswal, P. A. Hassan, A. G.
Wagh, Phys. B 2007, 398, 112; e) T. Chakraborty, I. Chakraborty, S. P.
Moulik, S. Ghosh, Langmuir 2009, 25, 3062.
[16] a) J. Aizenberg, A. Tkachenko, S. Weiner, L. Addadi, G. Hendler,
Nature 2001, 412, 819; b) B. Pokroy, V. Demensky, E. Zolotoyabko,
Adv. Funct. Mater. 2009, 19, 1054.
[17] a) Z. Tang, N. A. Kotov, S. Magonov, B. Ozturk, Nat. Mater. 2003,
2, 413; b) L. J. Bonderer, A. R. Studart, L. J. Gauckler, Science 2008,
319, 1069.
[18] a) P. Podsiadlo, A. K. Kaushik, E. M. Arruda, A. M. Waas, B. S. Shim,
J. Xu, H. Nandivada, B. G. Pumplin, J. Lahann, A. Ramamoorthy, N.
A. Kotov, Science 2007, 318, 80; b) M. A. Priolo, D. Gamboa, J. C.
Grunlan, ACS Appl. Mater. Interfaces 2009, 2, 312.
[19] G. Dewith, H. J. A. Vandijk, N. Hattu, K. Prijs, J. Mater. Sci. 1981, 16,
1592.
[20] W. C. Oliver, G. M. Pharr, J. Mater. Res. 1992, 7, 1564.
[21] a) S. Bec, A. Tonck, J. M. Georges, E. Georges, J. L. Loubet, Philos.
Mag. A-Phys. Condens. Matter Struct. Defect Mech. Prop. 1996, 74,
1061; b) S. Roche, S. Bec, J. L. Loubet, in Mechanical Properties
Derived from Nanostructuring Materials, Vol. 778 (Eds: D. F. Bahr,
H. Kung, N. R. Moody, K. J. Wahl), Materials Research Society,
Warrendale PA 2003, p. 117; c) G. Hochstetter, A. Jimenez, J. P.
Cano, E. Felder, Tribol. Int. 2003, 36, 973.
[22] J. Y. Rho, T. Y. Tsui, G. M. Pharr, Biomaterials 1997, 18, 1325.
[23] M. Antonietti, M. Breulmann, C. G. Göltners, H. Cölfen, K. K. W.
Wong, D. Walsh, S. Mann, Chem. Eur. J. 1998, 4, 2493.
[4] a) J. D. Currey, Proc. R. Soc. London, Ser. B 1977, 196, 443; b) A.
P. Jackson, J. F. V. Vincent, R. M. Turner, J. Mater. Sci. 1990, 25,
3173.
[5] C. Burger, H. W. Zhou, H. Wang, I. Sics, B. S. Hsiao, B. Chu,
L. Graham, M. J. Glimcher, Biophys. J. 2008, 95, 1985.
[6] S. Weiner, H. D. Wagner, Annu. Rev. Mater. Sci. 1998, 28, 271.
[7] D. Liu, H. D. Wagner, S. Weiner, J. Mater. Sci. Mater. Med. 2000, 11,
49.
[8] F. C. Meldrum, H. Cölfen, Chem. Rev. 2008, 108, 4332.
[9] a) G. Falini, S. Albeck, S. Weiner, L. Addadi, Science 1996, 271,
67; b) R. Kniep, S. Busch, Angew. Chem. Int. Ed. 1996, 35, 2624;
c) J. Aizenberg, A. J. Black, G. M. Whitesides, Nature 1999, 398, 495;
d) S. Sadasivan, D. Khushalani, S. Mann, Chem. Mater. 2005, 17,
2765.
[10] a) S. Mann, Nat. Mater. 2009, 8, 781; b) R.-Q. Song, H. Cölfen, Adv.
Mater. 2010, 22, 1301.
[11] H. Gao, B. Ji, I. L. Jäger, E. Arzt, P. Fratzl, Proc. Natl. Acad. Sci. U. S. A.
2003, 100, 5597.
[12] J. Xie, J. Y. Lee, D. I. C. Wang, J. Phy. Chem. C 2007, 111, 10226.
[13] C. E. Fowler, M. Li, S. Mann, H. C. Margolis, J. Mater. Chem. 2005,
15, 3317.
[14] a) L. Qi, J. Li, J. Ma, Adv. Mater. 2002, 14, 300; b) A. Kotachi,
T. Miura, H. Imai, Chem. Mater. 2004, 16, 3191.
[15] a) N. J. Turro, X. Lei, K. P. Ananthapadmanabhan, M. Aronson,
Langmuir 1995, 11, 2525; b) C. K. Ober, G. Wegner, Adv. Mater.
1997, 9, 17; c) S. De, A. Girigoswami, S. Das, J. Colloid Interface Sci.
Controlled formation of calcium-phosphate-based hybrid mesocrystals by
organic–inorganic co-assembly
Halei Zhai,a
Xiaobin Chu,a
Li Li,a
Xurong Xuab
and Ruikang Tang*ab
Received 28th July 2010, Accepted 27th August 2010
DOI: 10.1039/c0nr00542h
An understanding of controlled formation of biomimetic mesocrystals is of great importance in
materials chemistry and engineering. Here we report that organic–inorganic hybrid plates and even
mesocrystals can be conveniently synthesized using a one-pot reaction in a mixed system of protein
(bovine serum albumin (BSA)), surfactant (sodium bis(2-ethylhexyl) sulfosuccinate (AOT)) and
supersaturated calcium phosphate solution. The morphologies of calcium-phosphate-based products
are analogous to the general inorganic crystals but they have abnormal and interesting substructures.
The hybrids are constructed by the alternate stacking of organic layer (thickness of 1.31 nm) and
well-crystallized inorganic mineral layer (thickness of 2.13 nm) at the nanoscale. Their morphologies
(spindle, rhomboid and round) and sizes (200 nm–2 mm) can be tuned gradually by changing BSA,
AOT and calcium phosphate concentrations. This modulation effect can be explained by a competition
between the anisotropic and isotropic assembly of the ultrathin plate-like units. The anisotropic
assembly confers mesocrystal characteristics on the hybrids while the round ones are the results of
isotropic assembly. However, the basic lamellar organic–inorganic substructure remains unchanged
during the hybrid formation, which is a key factor to ensure the self-assembly from molecule to
micrometre scale. A morphological ternary diagram of BSA–AOT–calcium phosphate is used to
describe this controlled formation process, providing a feasible strategy to prepare the required
materials. This study highlights the cooperative effect of macromolecule (frame structure), small
biomolecule (binding sites) and mineral phase (main component) on the generation and regulation of
biomimetic hybrid mesocrystals.
Introduction
Scientists are eager to mimic nature’s ability to design functional
materials whose properties are often superior to the synthetic
ones. In nature, biominerals are widely produced by bacteria,
protists, plants, invertebrates and vertebrates, including
humankind.1
These biological materials are featured by a smart
combination of multi-components especially in the form of
integrated organic–inorganic hybrid materials, in which the
organic parts are often proteins and low-molecular-mass mole-
cules.2
They are constructed by using organic components to
control the nucleation, growth, organization and transformation
of inorganic phases. Interactions between organic and inorganic
phases at the molecular level, although complex, are common
occurrences to determine the size, shape, and properties of the
resulting products.1,3
Different from the synthesized ones, the
functions of biominerals depend to a large extent on the ordered
association of biomolecules with mineral phases. The organized
hybrid materials, unlike the single components, can be tailored
into different compositions and morphologies, e.g. bone,4
tooth5
and mollusc shells6
etc., to ensure the optimal mechanical and
physicochemical characteristics.
The controls that determine the sizes, shapes, and properties of
crystals are a key to addressing numerous challenges in material
designs and applications. It has been revealed that organic
molecules can influence the shape and properties of inorganic
crystals.7
However, it is difficult for the two distinct organic and
inorganic phases to spontaneously assemble into highly ordered
structures. In living organisms, biological mineralization is able
to combine particular building blocks or entities into functional
hybrid composites. An understanding of these biochemical
controls is essential and important, not only to study
biomineralization mechanisms further, but also to design novel
hybrid materials and processing technologies. Despite the
complicated hierarchical structures of biominerals, their basic
building blocks are frequently the nano-sized organic–inorganic
composites.8
Therefore, an ordered and periodic assembly of
organic and inorganic nanophases at the nanoscale is crucial to
biomimetically synthesize hybrid materials. But, how can we
design ordered hybrid composites and how can we conveniently
control their structures, sizes and morphologies under mild
conditions?
Although organic–inorganic hybrid materials have been
approached by various methods such as layer-by-layer (LbL)9
and template-directed crystallization,10
the bottom-up fabrica-
tion from ions or molecules is still a great challenge in the
laboratory since the control of periodic deposition is difficult to
achieve at hierarchical scales. In conventional biomimetic crys-
tallization studies, organic molecules, which act as structure-
directing agents, modulate the crystal morphology by their
a
Centre for Biomaterials and Biopathways, Zhejiang University,
Hangzhou, Zhejiang, 310027, China. E-mail: rtang@zju.edu.cn; Fax:
+86-571-87953736; Tel: +86-571-87953736
b
State Key Laboratory of Silicon Materials, Zhejiang University,
Hangzhou, Zhejiang, 310027, China
2456 | Nanoscale, 2010, 2, 2456–2462 This journal is ª The Royal Society of Chemistry 2010
PAPER www.rsc.org/nanoscale | Nanoscale
Publishedon13October2010.Downloadedon24/01/201605:58:33. View Article Online / Journal Homepage / Table of Contents for this issue
selective absorption onto crystal faces, altering crystal facet
stability and growth kinetics.7,11
Recently, a non-classical crystal
growth pathway based upon nano assembly has received
considerable attention.12
The nanoparticles, which are directed
by specific organic additives, can act as the basic building units to
assemble into superstructures or mesocrystals. During such
a process, the organic molecules (especially macromolecules)
selectively absorb and interact with primary nanocrystals. The
assembly process follows programmed arrangement into high
order hybrid structures.13
The morphology can be tuned by
varying the interactions between different organic and inorganic
phases. However, the one step bottom-up process, which starts
from the molecular level rather than from preformed nano-
particle precursors, may be readily able to control the orientation
and order of assembly processes to form integrated hybrid
nanocomposite. But this strategy requires a precise and sponta-
neous co-assembly of both organic and inorganic phases
alternately at both the molecular level and the nanoscale.14
In this paper, we reported an easy but effective method for
direct synthesis of organic–inorganic hybrid mesocrystals by
a emergent co-assembly process of protein (bovine serum
albumin (BSA)) and surfactant (sodium bis(2-ethylhexyl) sulfo-
succinate (AOT)) in a supersaturated calcium phosphate solu-
tion. The calcium-phosphate-based hybrid crystals with lamellar
structure have different properties from conventional ones. Here
we emphasize that the size and morphology of the resulting
hybrids could be regulated readily by varying BSA, AOT and
calcium phosphate concentrations according to a suggested
morphological ternary diagram. This study provided a novel
pathway to one-pot preparation of functional hybrid crystal
materials with tuneable size and morphologies by organic–inor-
ganic co-assembly.
Results and discussion
It is believed that functional organic molecules can interact with
calcium species at the organic–inorganic interfaces to modulate
the growth and assemble of inorganic crystals. BSA is one of the
most studied proteins but this biological macromolecule is not an
effective modifier in calcium phosphate crystallization.15
It has
been previously confirmed that the interaction between BSA and
calcium or phosphate ions in aqueous solutions is poor.16
BSA
itself is inert in mineral deposition. In contrast, many surfactant
molecules are widely used as effective promoters and templates in
biomimetic calcium mineralization since their hydrophilic groups
(especially the sulfonate and carboxylate groups) provide active
binding sites to calcium ions. AOT is one among typical agents
that can modulate calcium phosphate precipitation significantly.
AOT molecules have a strong binding effect with calcium ions
due to their highly charged -SO3
2À
groups.16,17
However, hierar-
chical or complicated biomineral-like structures cannot be
achieved by using this small molecule due to the lack of higher-
order structures. In our control experiments, only poor crystal-
line HAP was obtained if BSA was added into the supersaturated
calcium phosphate solutions; AOT alone produced the conven-
tional rod-like HAP crystals without any organized hybrid
structure. These results matched the previous studies and
understandings well. However, the cooperative effect of BSA and
AOT in the calcium phosphate solution could lead to the
formation of unique hybrids in a one-pot reaction.
Under an experimental condition of 2 mM AOT, 1 mg mlÀ1
BSA and 1.25 mM calcium ions (the molar ratio of calcium to
phosphate was fixed at 1.67 in all experiments), the uniform
rhombic plates precipitated spontaneously as shown by scanning
electron microscopy (SEM, Fig. 1(A)). Their size distribution
was homogeneous. The typical rhombic plates were 1.23 Æ 0.21
and 0.91 Æ 0.18 mm along their long and short axes, respectively
(statistical results from $100 plates); the aspect ratio was about
1.4. The thickness of the plates was 130 Æ 20 nm. These rhombic
plates had exactly same morphology (Fig. 1(B)) and this char-
acteristic was similar to the general inorganic crystals. However,
the chemical compositions of the obtained plates were relatively
complicated. Besides the elements of calcium and phosphorus,
the element of sulfur was detected in the solids by using energy-
dispersive X-ray spectroscopy (EDS). This result indicated the
presence of AOT (-SO3
2À
) in the hybrid plates. It was also
revealed that inorganic part in the plates was a kind of calcium
phosphate minerals with Ca : P molar ratio of 1.5–1.6. The
coexistence of organic–inorganic components was also
confirmed by Fourier transform infrared spectroscopy (FT-IR,
Fig. 1(C)). The peaks at 1737, 1459 and 1419 cmÀ1
were the
characteristic signals of AOT, while the bands at 1656 (amide I)
and 1555 cmÀ1
(amide II) showed the involvement of BSA in the
solids.18
The broad peaks at 1022 and 564 cmÀ1
were assigned to
the inorganic phosphate groups.19
Thermogravimetric analysis
(TGA) showed that the mineral phase was the main composition
in the solids. The weight loss of 38% between 100 and 500 
C was
corresponded predominantly to removal of the organic phase,
while the weight contents of the inorganic phases were 62%. In
addition, the plates became ‘crimped-paper’-like after calcina-
tions at 500 
C in air for 2 h. Without the organic frame, the
solids became brittle and the structures were collapsed readily
into small pieces under an ultrasonic condition. Many previous
studies suggested that the organic compounds play a regulation
role in inorganic mineralization rather than being involved in
Fig. 1 (A) SEM image of the rhombic plates. (B) Enlarged image of the
rhombic plate in the white circle; the double-headed arrow shows the
extended orientation. (C) FT-IR pattern of the products. (D) The
rhombic plates after calcination at 500 
C in air.
This journal is ª The Royal Society of Chemistry 2010 Nanoscale, 2010, 2, 2456–2462 | 2457
Publishedon13October2010.Downloadedon24/01/201605:58:33. View Article Online
structural recombination. However, the current results implied
that BSA and AOT were the key components in the hybrid
construction. Thus, these solids were different from the other
precipitated inorganic crystals in the presence of organic addi-
tives.
The resulting rhombic plates shared the same size and aniso-
tropic morphology similar as general inorganic crystals.
However, in-depth examination revealed that they were distinct
from the conventional calcium phosphate crystals.20
The
rhombic plates were examined by wide angle X-ray diffraction
(WAXD, Fig. 2(A)) and as expectated, the crystalline HAP-like
calcium phosphate phase was detected. The WAXD pattern was
very similar to that of pure HAP but small peak shifts were also
observed. We suggested that the binding effect between the
organic component and calcium ions would cause the lattice
distortion. The lattice structure of the inorganic phase could be
revealed at the atomic scale by using high resolution transmission
electron microscopy from a top view of the plates (HRTEM,
Fig. 2(B)). This image represents a typical ultrathin inorganic
crystal layer embedded in the rhombic plates. However, another
independent set of diffraction peaks was found in the X-ray
diffraction (XRD) pattern, which revealed that a superstructure
was present in the hybrids. The characteristic peaks of lamellar
structure (interspacing distance, d ¼ 3.43 nm) could be found
from both small angle X-ray diffraction (SAXD) and WAXD
(Fig. 2(A)), indicating an ordered arrangement of subunits along
a crystallographic direction rather than a simple mixture of the
organic and inorganic phases. A side view of the ultra-thin
sectioned samples under transmission electron microscopy
(TEM, Fig. 2(C)) confirmed the internal structure: the organic
layers (light, 1.31 nm) and the inorganic layers (dark, 2.13 nm)
alternately stacked at the nanoscale to form the compact hybrid
structure. Thus, the organic molecules (BSA and AOT) were well
organized to form the layered organic phase. Each organic–
inorganic ultra-thin unit had a thickness of 3.44 nm, which
agreed with the XRD data, 3.43 nm, and the individual inorganic
layer was a calcium phosphate crystal plate with a thickness of
only 2.13 nm. These nanoplates acted as the building blocks that
could self-assemble together with the organic layers to generate
the lamellar complex. Additionally, a wave-like superficial
texture of the hybrids could be observed (Fig. 1(B)) and the
profiles were similar to the hybrid crystal morphology. This
phenomenon indicated that the assembly might be an anisotropic
process.
In order to understand the orientation of each inorganic layer,
selected area electron diffraction study (SAED, Fig. 2(D)) was
applied. It was noted that the anisotropic diffraction dots rather
than the isotropic diffraction rings were obtained during the
examination of a whole rhombic crystal, which represented
a similar characteristic of single crystal. It was interesting that the
orientation reflected by these dots (arrow in the insert image) was
exactly same as the long axis of the examined rhombic crystal.
Such a coincidence implied that all the ultrathin inorganic crystal
layers within the hybrid plates should share the same crystallo-
graphic orientation. Additionally, the experimental diffractions
dots of the whole crystal were almost same as the fast Fourier
transform (FFT) result (Fig. 2(B)) of an individual crystal layer.
Therefore, the formed hybrid crystal exhibited similar features to
a single crystal; however, it had additional superlattice structure.
Since the rhombic plates had a specific morphology while they
were not constructed as the conventional single crystals, these
hybrids could be considered as a kind of artificial meso-
crystal.12,21
However, the imperfect dots on Fig. 2 (D) might
indicate that the misaligned orientation still occurred during
nano assembly. Since the material was constructed by ultrathin
calcium phosphate units, it was interesting that flexible and
elastic features were conferred onto the mesocrystal along the
lamellar packing direction in spite of that; its main composition
was a brittle ceramic phase. These mechanical properties of the
hybrids had been characterized by our previous study,16
demonstrating the advantages of organized assembly for
formation of mesocrystals in material functionalization.
The convenient control of the size and morphology of the
organic–inorganic hybrids and mesocrystals is a challenge,
although those for single hybrid crystals are nowadays sophis-
ticated. In our experiments, the calcium phosphate–BSA–AOT
hybrid mesocrystals with different size and morphology could be
feasibly regulated within a simple reaction system by changing
the reactant concentrations. We fixed BSA and calcium
concentration at 0.5 mg mlÀ1
and 1.25 mM, respectively. When
the AOT concentration was 1.00 mM, the obtained hybrid plates
were not rhombic plates any more. Their shapes became spindle-
like. The hybrid plates changed into a round shape when the
AOT concentration was increased to 4.00 mM. However, the
further decreasing or increasing of AOT concentration result into
the disappearance of the co-assembly or hybrid in the system. In
this experiment, their morphologies were gradually adjustable
from spindle, to rhombus to round by increasing the AOT
concentration from 1.00 to 4.00 mM (Fig. 3). During the
evolution process, the length along the short axis of the formed
Fig. 2 (A) WAXD and SAXD (insert) patterns of the rhombi;
(B) HRTEM of a rhombus (top view). Insert: FFT simulation result;
(C) TEM image of ultra-thin sectioned rhomb from side view. The values
of 2.13, 1.31 and 3.44 nm corresponded to the thicknesses of inorganic
(dark), organic (light) and organic–inorganic complex layers, respec-
tively. Insert: TEM image of the side view of the ultra-thin sections of the
plates, bar is 0.5mm. (C) is the enlargement of the region within the white
circle; (D) TEM image of the hybrids. Insert was the SEAD pattern
(white circle area). The HRTEM image in (B) was also obtained on the
same area by the in situ technique. Arrows showed that each individual
inorganic plate in the hybrid shared the same crystallographic orienta-
tion, which was the long axis of the rhombus.
2458 | Nanoscale, 2010, 2, 2456–2462 This journal is ª The Royal Society of Chemistry 2010
Publishedon13October2010.Downloadedon24/01/201605:58:33. View Article Online
hybrid plates did not change significantly, it was maintained at
300–400 nm. However, the long axis kept on decreasing from
1.50 mm to 300–400 nm with increasing AOT concentration.
Accordingly, the hybrid morphology became isotropic. This
phenomenon implied that AOT component was an important
factor to control the a degree of anisotropic co-assembly of the
hybrids.
Although the morphologies and sizes of the resulted hybrid
plates were influenced remarkably by the changing of AOT
concentrations in the reaction solutions, the internal organic–
inorganic subunit remained. The WAXD and SAXD patterns of
the spindles, rhombi and rounds were exactly same without any
change. But the misalignments of each individual inorganic layer
in the hybrid increased with the increasing of AOT concentra-
tion. The crystallographic mismatch of the inorganic layers could
be examined by using SEAD. During the evolution from the
regular rhombi to round shapes, the diffraction dots disappeared
gradually while the diffraction rings existed (Fig. 4). This
tendency indicated that the preferred orientation of the thin
calcium phosphate planes in the hybrid was weakened. Although
AOT itself could result in aggregates in solution to induce
calcium mineralization, the aggregation was simple and isotropic
due to the lack of complicated configuration. Therefore, it was
reasonable that the excessive AOT could destroy the anisotropic
assembly of the ultrathin mineral plates in the hybrid rhombi.
Although the inorganic and organic layers were still packed
layer-by-layer strictly along the thickness direction, the crystal-
lographic directions of the inorganic crystal planes in the hybrids
became disordered. The anisotropic assembly transformed the
orientation of the long axes into the isotropic mode with
increasing AOT concentration; thus, the round plates were
finally yielded at 4.00 mM AOT and the hybrid was not meso-
crystalline any more. Besides, it should be mentioned that the
percentages of organic and inorganic contents in the hybrid
solids was not changed significantly during the morphology
modulation; in which the inorganic content was kept within
a range of 69–72% from the spindles to the rounds.
Besides the AOT concentrations, the formation of hybrid
crystals could be also adjustable by BSA concentration. In this
examination, the concentrations of AOT and calcium were
maintained at 2 mM and 1.25 mM, respectively, and the BSA
concentrations were increased from 0.25 to 2.00 mg mlÀ1
. It was
noted that the morphologies of hybrid plates underwent another
gradual evolution from the irregular quadrilaterals to rhombi
and then to plump spindles (Fig. 5). The sizes and aspect ratios of
the hybrids increased from 200 nm to 2 mm and 1.1 to 2.0,
respectively, during the modulation. Although the hybrid width
increased along the short axis, the more extended length along
the long axis indicated that the anisotropy assembly process was
affected significantly by the protein concentration. It was noticed
that in biomineralization, the complicated hierarchical building
structures of biominerals are frequently contributed by the
ordered aggregates of proteins. Again, the basic organic–inor-
ganic units and their ordered packing behaviours were not
changed during the morphology and size regulations. It was
mentioned that, when the BSA concentration increased, the role
of AOT in the synthesis decreased. Therefore, the ratio or the
Fig. 3 SEM images of the hybrids synthesised at AOT concentrations of
1.00 (A), 2.00 (B) and 4.00 mM (C). (D)–(F) are the corresponding XRD
patterns of (A)–(C), respectively.
Fig. 4 During the morphology change from rhombus (A) to round (B),
anisotropic diffraction dots became isotropic rings in the corresponding
SEAD pattern.
Fig. 5 SEM images of the hybrids at BSA concentration of 0.25 (A),
1.13 (B) and 2.00 mg mlÀ1
(C).
This journal is ª The Royal Society of Chemistry 2010 Nanoscale, 2010, 2, 2456–2462 | 2459
Publishedon13October2010.Downloadedon24/01/201605:58:33. View Article Online
cooperative effect of BSA and AOT was another key factor in
mesocrystal formation and regulation.
It was known that the co-assembly could not occur in the
absence of the inorganic phase. Thereby, it was reasonable that
the concentrations of calcium and phosphate could control the
mesocrystal formation too (Fig. 6). Under BSA and AOT
concentrations of 1.00 mg mlÀ1
and 2.00 mM, respectively, the
resulting rhombi shared the same intermediate state with
increasing calcium and phosphate concentration in the reaction
solution. If calcium concentration was decreased to 0.63 mM, the
poly-dispersed quadrilaterals-like plates (size of 400–800 nm)
formed with the small aspect ratio of 1.1. If the concentration
was increased to 2.50 mM, the slender spindle-like plates were
obtained and their size distribution was 1.8–2.3 mm with an
aspect ratio was 2.5. From the evolution from quadrilaterals,
rhombi to slender spindles, it could be seen that the anisotropic
co-assembly process was enhanced.
The previous studies of biomimetic fabrication of hybrid
materials with artificial molecules such as peptide-amphiphile,22
block copolymer,23
and amphiphilic dendro-calixarene,24
sug-
gested that the specific sites and sterically constrained effect may
control the assembly of the organic template and then the size
and morphology of the final hybrid materials. Different from the
above-mentioned understanding, under our experimental
conditions, the change of BSA and AOT concentrations were
directly related to the different modification state of BSA. The
BSA protein, which was constituted by a single chain of
583 amino acid residues, acted as a stable and relatively rigid
fragment connected with the special motif (AOT aggregates).25
The hydrophilic groups of aggregates exposed to aqueous solu-
tions and their configuration can be adjusted. The highly charged
group (–SO3
2À
) in AOT could greatly interact with calcium ions
and then modulated calcium phosphate precipitation
significantly, which had been demonstrated experimentally in
many works and in our previous paper.16,26
However, the binding
ability of BSA with calcium ions is weak and the controlling
effect on the mineral formation is relatively poor. As a result,
BSA acted as structural frame while the AOT aggregates
provided the nucleation sites of mineral during the co-assembly
process. In the current study, BSA macromolcules combined
with smaller AOT molecules to form a BSA–AOT complex and
such a modified protein could effective control the crysallization
and assembly of the calcium phosphate mineral. To some extent,
this method provides an efficient way to turn a non-mineraliza-
tion protein into a mineralization protein by using surfactants.
The conformation of the macromolecules restricts the assembly
only along certain specific directions. However, the larger
concentration of AOT is accompanied by an increase in the
amount and size of AOT aggregates, offering more sites for
the assembly process.27
As a result, the controlling effect from the
protein was counteracted and the assembly process could happen
at more directions to form the isotropic rounds. Furthermore,
increasing the amount or the relative amount of BSA concen-
trations partly restricted the assembly process in specific prefer-
ential orientations by spatial configuration to form the
anisotropic hybrids or mesocrystals.28
Thus, the co-assembly
process preferred to occur in certain directions, especially along
the long axis of the hybrid plates rather than the short axis.
Although the short axis partly extended under some experi-
mental cases, the greatly increase along the long axis resulted into
the spindles-like mesocrystal formation. The competitive
controlling effect of BSA and AOT led to the transformation of
an isotropic and anisotropic assembly process during hybrid
crystal construction. Thus, the formation of different hybrids
and mesocrystals with tuneable size and morphologies could be
achieved.
An anisotropic co-assembly process could also be promoted by
increasing the mineral ion concentrations. In the formation
process of mesocrystals, the inorganic precursor controlled the
size and morphology of the final product by tuning the amounts,
size and shapes of the nano-sized building blocks.29
Under our
experimental conditions, the controlling role of mesocrystal
growth became dominant in greater saturation to decide the
product size and structure. As the preferred orientation of
the calcium phosphate crystal plates is parallel to the long axis of
the rhombic plates, the fast growth of the calcium phosphate
plate crystals along this preferred orientation promoted the
formation of the slender spindle-like plates with larger aspect
ratios during the co-assembly process. However, the interaction
between BSA-AOT complex and calcium phosphate crystal was
also responsible for the co-assembly of the organic and inorganic
phase to form highly ordered hybrid materials and maintain their
internal structure.
Actually, the generation of hybrid material via the cooperative
effect of macromolecules (mainly proteins), small biomolecules
and the mineral phase is a common strategy in natural bio-
mineralization.30
In the biological construction, high-molecular-
weight macromolecules, such as collagen, act as support matrix
to provide a structural frame for the mineralization, the
biomineralization proteins themselves have nucleation sites
but most matrices receive mineralization function by binding
and stabilizing functional motifs that are carboxylate- or
Fig. 6 SEM images of the hybrids at calcium concentration of 0.63 (A),
1.56 (B) and 2.50 mM. In all experiments, the ratio of calcium to phos-
phate in the reaction solution was maintained at 1.67.
2460 | Nanoscale, 2010, 2, 2456–2462 This journal is ª The Royal Society of Chemistry 2010
Publishedon13October2010.Downloadedon24/01/201605:58:33. View Article Online
sulfonate-rich. Thus, the combination of organic–inorganic
mineralization interfaces and the organized organic matrices can
concentrate the mineral ions to induce the deposition as well as to
regulate the size, morphology and orientation of the inorganic
building blocks to form integrated organic–inorganic hybrid
composites with complicated structure. We suggest that in this
system, BSA is the structural frame to control the anisotropic
assembly; the adsorption of AOT onto BSA enhances the
mineralization ability of the protein; and the mineral acts as an
inorganic conjunction phase to solidify the organic–inorganic
hybrid structure. In the experiments, the increase of BSA
promoted the formation of larger hybrid plates with increased
aspect ratio, while AOT exhibited the opposite controlling effect.
The increasing of inorganic concentrations preferred the
formation of slender hybrid plates with a larger size. In order to
show the controlling effect of the reactant concentrations, the
simplified morphological maps in the form of solution ternary
diagrams was proposed (Fig. 7). The biomimetic formation
hybrid and mesocrystals could be yielded in the grey region. In
the specific regions, the formed hybrid plates had a similar size
and morphology. From points A to B, the increase of aspect ratio
was preferred as the hybrid rounds transformed into the spindle
ones. Since the anisotropic assembly behaviour was enhanced,
this evolution implied that the resulting mesocrystals became
more organized and the mismatch degrees of the inorganic layers
in the hybrids could be reduced. From points A to C, both the
size and aspect ratio of the resulted hybrids were increased and
their morphologies were changed from rounds to spindles. From
points B to C, the hybrids turned from wide spindles to slender
spindles with increased size and aspect ratio too. By using this
morphological ternary diagram, we could design readily hybrids
and mesocrystals with the required size and morphology.
Conclusions
We demonstrate that the ordered and uniform hybrids or mes-
ocrystals can be biomimetically synthesized by the co-assembly
of proteins, small functional molecules and minerals using
a simple one-pot reaction. Their size distributions and
morphologies can be adjusted by varying the component
concentration in reaction solutions. The anisotropic co-assembly
of the BSA–AOT complex and ultrathin calcium phosphate
crystal plates is a key to the control of mesocrystal formation.
A morphological ternary diagram can be used to design different
hybrid materials as requireed. This work may give another
inspiration to the assembly of multi components into one inte-
grated hybrid material with a highly ordered structure.
Furthermore, the bottom-up pathway of controlled fabrication
may be developed as a simple and effective strategy to prepare
feasibly functional hybrid and mesocrystal materials.
Experimental
Materials
Triply distilled water was used in all the experiments. Ca(NO3)2
and (NH4)2HPO4 were of analytical and their solution were
filtered twice using 0.22mm Millipore films prior to use. BSA
(Albumin Bovine fraction V, BR, purity  98%, LABMAX) and
AOT (Aldrich) were used without any further purification.
Hybrid plate preparation
Using a typical experiment as an example, 100 ml aqueous
solution containing 4 mM AOT and 0.20 g BSA was mixed with
50 ml Ca(NO3)2 solution (5mM). The solution pH was adjusted
to 10.0 Æ 0.5 at room temperature by 3 M ammonia solution.
Then 50 ml (NH4)2HPO4 solution (3 mM, pH ¼ 10.0 Æ 0.5) was
added dropwise at a rate of 1.5 ml minÀ1
. The reaction solution
contained 2.00 mM AOT, 1.00 mg mlÀ1
BSA, 1.25 mM Ca(NO3)2
and 0.75 mM (NH4)2HPO4. The mixture was gently stirred at
30 Æ 1 
C for 24 h. The precipitated solids were collected by
centrifugation at 6000 rpm. The solid were washed by water for
three times and were vacuum-dried at 35 Æ 1 
C. In order to
examine the controlling effect of reactant concentrations on
hybrid formation, different concentrations of AOT, BSA and
calcium phosphate ions were used and all the experimental
processes were the same.
Fig. 7 Controlled synthesis of hybrids by a morphological ternary diagram. The co-assembly occurred within the grey area and the formation of
mesocrystals was preferred in its left and bottom sections. The typical morphology of the final products were also demonstrated. Bar ¼ 1mm.
This journal is ª The Royal Society of Chemistry 2010 Nanoscale, 2010, 2, 2456–2462 | 2461
Publishedon13October2010.Downloadedon24/01/201605:58:33. View Article Online
Characterizations
SEM was performed by using a HITACHI S-4800 at a typical
acceleration voltage of 5 kV. FT-IR spectra (Nicolet Nexus 670)
were applied to analysis the hybrid compositions. WAXD and
SAXD were characterized by a Rigaku D/max-2550pc with
monochromatized Cu-Ka radiation; the scanning step was 0.02
.
TGA was performed by a TA Instrument SDT Q600. The
experiment was measured in a temperature range from room
temperature to 1000 
C under nitrogen atmosphere. TEM
observations were performed by a CM200UT TEM (Philips) at
an acceleration voltage of 160 kV. During the ultra-thin
sectioned TEM examination, rhombi were embedded in epoxy.
The mixture was solidified at 80 
C for 12 h and then carefully
microtomed by a Reichert-Jung Ultracut E using a diamond
knife.
Acknowledgements
We thank Jieru Wang, Xinting Cong, Xiaomin Tang, Yin Xu
and Linshen Chen for their help with characterization, Haihua
Pan and Yuan Su for discussions. This work was supported by
the Fundamental Research Funds for the Central Universities,
National Natural Science Foundation of China (20871102),
Zhejiang Provincial Natural Science Foundation (R407087) and
Daming Biomineralization Foundation.
Notes and references
1 S. Mann, Biomineralization: Principles and Concepts in Bioinorganic
Materials Chemistry, Oxford University Press, 2001.
2 L. Bedouet, F. Rusconi, M. Rousseau, D. Duplat, A. Marie,
L. Dubost, K. Le Ny, S. Berland, J. Peduzzi and E. Lopez, Comp.
Biochem. Physiol., Part B: Biochem. Mol. Biol., 2006, 144, 532–543;
J. L. Arias and M. a. S. Fernacndez, Chem. Rev., 2008, 108,
4475–4482.
3 C. E. Killian and F. H. Wilt, Chem. Rev., 2008, 108, 4463–4474;
J. S. Evans, Chem. Rev., 2008, 108, 4455–4462.
4 S. Weiner and H. D. Wagner, Annu. Rev. Mater. Sci., 1998, 28,
271–298.
5 S. Busch, U. Schwarz and R. Kniep, Chem. Mater., 2001, 13,
3260–3271.
6 N. Watabe, J. Ultrastruct. Res., 1965, 12, 351–370.
7 F. C. Meldrum and H. C€olfen, Chem. Rev., 2008, 108, 4332–4432.
8 R. Z. Wang, Z. Suo, A. G. Evans, N. Yao and I. A. Aksay, J. Mater.
Res., 2001, 16, 2485–2493; H. J. Gao, B. H. Ji, I. L. Jager, E. Arzt and
P. Fratzl, Proc. Natl. Acad. Sci. U. S. A., 2003, 100, 5597–5600.
9 Z. Tang, N. A. Kotov, S. Magonov and B. Ozturk, Nat. Mater., 2003,
2, 413–418; P. Podsiadlo, A. K. Kaushik, E. M. Arruda, A. M. Waas,
B. S. Shim, J. Xu, H. Nandivada, B. G. Pumplin, J. Lahann,
A. Ramamoorthy and N. A. Kotov, Science, 2007, 318, 80–83.
10 N. Gehrke, N. Nassif, N. Pinna, M. Antonietti, H. S. Gupta and
H. C€olfen, Chem. Mater., 2005, 17, 6514–6516; P. H. Kithva,
L. Grondahl, R. Kumar, D. Martin and M. Trau, Nanoscale, 2009,
1, 229–232.
11 N. A. J. M. Sommerdijk and G. d. With, Chem. Rev., 2008, 108,
4499–4550.
12 R. Q. Song and H. C€olfen, Adv. Mater., 2010, 22, 1301–1330.
13 M. Li, H. C€olfen and S. Mann, J. Mater. Chem., 2004, 14, 2269–2276.
14 S. Mann, Nat. Mater., 2009, 8, 781–792.
15 R. I. Martin and P. W. Brown, J. Mater. Sci.: Mater. Med., 1994, 5,
96–102; K. L. Yadav and P. W. Brown, J. Biomed. Mater. Res., 2003,
65a, 158–163.
16 H. Zhai, W. Jiang, J. Tao, S. Lin, X. Chu, X. Xu and R. Tang, Adv.
Mater., 2010, 22, 3729–3734.
17 C. E. Fowler, M. Li, S. Mann and H. C. Margolis, J. Mater. Chem.,
2005, 15, 3317–3325.
18 G. Falini, S. Weiner and L. Addadi, Calcif. Tissue Int., 2003, 72,
548–554.
19 S. J. Gadaleta, E. P. Paschalis, F. Betts, R. Mendelsohn and
A. L. Boskey, Calcif. Tissue Int., 1996, 58, 9–16.
20 J. Song, V. Malathong and C. R. Bertozzi, J. Am. Chem. Soc., 2005,
127, 3366–3372; A. Ethirajan, U. Ziener, A. Chuvilin, U. Kaiser,
H. C€olfen and K. Landfester, Adv. Funct. Mater., 2008, 18,
2221–2227; Y. Zhang and J. Lu, Cryst. Growth Des., 2008, 8,
2101–2107.
21 A.-W. Xu, M. Antonietti, S.-H. Yu and H. C€olfen, Adv. Mater., 2008,
20, 1333–1338.
22 J. D. Hartgerink, E. Beniash and S. I. Stupp, Science, 2001, 294,
1684–1688; V. M. Yuwono and J. D. Hartgerink, Langmuir, 2007,
23, 5033–5038.
23 Z. H. Nie, D. Fava, E. Kumacheva, S. Zou, G. C. Walker and
M. Rubinstein, Nat. Mater., 2007, 6, 609–614; H. Wang, A. J. Patil,
K. Liu, S. Petrov, S. Mann, M. A. Winnik and I. Manners, Adv.
Mater., 2009, 21, 1805–1808.
24 M. Kellermann, W. Bauer, A. Hirsch, B. Schade, K. Ludwig and
C. B€ottcher, Angew. Chem., Int. Ed., 2004, 43, 2959–2962.
25 N. J. Turro, X. G. Lei, K. P. Ananthapadmanabhan and M. Aronson,
Langmuir, 1995, 11, 2525–2533; S. De, A. Girigoswami and S. Das,
J. Colloid Interface Sci., 2005, 285, 562–573.
26 S. Sarda, M. Heughebaert and A. Lebugle, Chem. Mater., 1999, 11,
2722–2727.
27 C. K. Ober and G. Wegner, Adv. Mater., 1997, 9, 17–31.
28 H.-A. Klok, J. F. Langenwalter and S. Lecommandoux,
Macromolecules, 2000, 33, 7819–7826; X. Kong and S. A. Jenekhe,
Macromolecules, 2004, 37, 8180–8183; L. Rubatat, X. Kong,
S. A. Jenekhe, J. Ruokolainen, M. Hojeij and R. Mezzenga,
Macromolecules, 2008, 41, 1846–1852; A. Sacnchez-Ferrer and
R. Mezzenga, Macromolecules, 2010, 43, 1093–1100.
29 H. C€olfen and M. Antonietti, Angew. Chem., Int. Ed., 2005, 44,
5576–5591.
30 N. Kroger, R. Deutzmann, C. Bergsdorf and M. Sumper, Proc. Natl.
Acad. Sci. U. S. A., 2000, 97, 14133–14138; L. C. Palmer,
C. J. Newcomb, S. R. Kaltz, E. D. Spoerke and S. I. Stupp, Chem.
Rev., 2008, 108, 4754–4783.
2462 | Nanoscale, 2010, 2, 2456–2462 This journal is ª The Royal Society of Chemistry 2010
Publishedon13October2010.Downloadedon24/01/201605:58:33. View Article Online
Spontaneously amplified homochiral organic–inorganic
nano-helix complexes via self-proliferation†
Halei Zhai,a
Yan Quan,a
Li Li,a
Xiang-Yang Liu,b
Xurong Xuc
and Ruikang Tang*ac
Most spiral coiled biomaterials in nature, such as gastropod shells, are homochiral, and the favoured chiral
feature can be precisely inherited. This inspired us that selected material structures, including chirality,
could be specifically replicated into the self-similar populations; however, a physicochemical
understanding of the material-based heritage is unknown. We study the homochirality by using calcium
phosphate mineralization in the presence of racemic amphiphilic molecules and biological protein. The
organic–inorganic hybrid materials with spiral coiling characteristics are produced at the nanoscale. The
resulted helixes are chiral with the left- and right-handed characteristics, which are agglomerated
hierarchically to from clusters and networks. It is interesting that each cluster or network is homochiral
so that the enantiomorphs can be separated readily. Actually, each homochiral architecture is evolved
from an original chiral helix, demonstrating the heritage of the matrix chirality during the material
proliferation under a racemic condition. By using the Ginzburg–Landaue expression we find that the
chiral recognition in the organic–inorganic hybrid formation may be determined by a spontaneous
chiral separation and immobilization of asymmetric amphiphilic molecules on the mineral surface, which
transferred the structural information from the mother matrix to the descendants by an energetic
control. This study shows how biomolecules guide the selective amplification of chiral materials via
spontaneous self-replication. Such a strategy can be applied generally in the design and production of
artificial materials with self-similar structure characteristics.
1 Introduction
Through long time periods of evolution, most spiral coiled bio-
materials in nature, like gastropod shells, adopt a specic
homochirality.1,2
For example, the majority of current gastropod
shells have a right-handed (R-) coiling pattern (Fig. 1a).3
The
chiral minority was eliminated eventually by a frequency-
dependent selection and the dominant one proliferated.4,5
This
inspires us that materials with specic structural properties,
like chirality, can spontaneously develop into a large self-similar
community.6
Biologically, it is accepted that regularly expressed
biomolecules, together with inorganic minerals, constitute the
physical chirality of gastropod offsprings under the guidance of
a controlling gene.7
For instance, at the growth front of shells,
the tiny chitin nanocrystals behave as the amphiphilic mole-
cules and self-assemble into the liquid crystal layers (Fig. 1b).8
Fig. 1 The chirality of gastropod shells and a schematic drawing of the shell
mineralization front. (a) General gastropod species have right-handed shells. (b)
During the natural generation of shell structure b-chitin molecules assemble into
supermolecules (chitin crystallites) and their liquid-crystal layers induce the spiral
mineralization of calcium carbonate (this scheme is prepared based upon a
mechanism proposed by Cartwright et al.).8,9
a
Centre for Biomaterials and Biopathways and Department of Chemistry, Zhejiang
University, Hangzhou, 310027, China. E-mail: rtang@zju.edu.cn; Fax: +86 571-
8795-3736
b
Department of Physics and Department of Chemistry, National University of
Singapore, Singapore 117542, Singapore
c
Qiushi Academy for Advanced Studies, Zhejiang University, Hangzhou, Zhejiang
310027, China
† Electronic supplementary information (ESI) available: Supporting gures and
tables. See DOI: 10.1039/c3nr33782k
Cite this: Nanoscale, 2013, 5, 3006
Received 23rd November 2012
Accepted 29th January 2013
DOI: 10.1039/c3nr33782k
www.rsc.org/nanoscale
3006 | Nanoscale, 2013, 5, 3006–3012 This journal is ª The Royal Society of Chemistry 2013
Nanoscale
PAPER
Publishedon06February2013.DownloadedbyUniversityofCalifornia-SanFranciscoon24/01/201605:50:31.
View Article Online
View Journal | View Issue
These chitin layers provide growing sites for the inorganic
phase and modulate the mineralization together with related
proteins. Thus, it follows that the spiral micro-pattern consti-
tuted by a chitin–calcium carbonate lamellar structure is grad-
ually constructed (Fig. 1b).9
In this sense, an understanding on
the physicochemical regulations of the organic–inorganic bio-
inspired materials with selective chirality will advance our
knowledge in chemistry and materials sciences. A challenging
question to be addressed is whether we can mimic the self-
evolution (symmetry breaking) process of shells in our labora-
tories so that the chiral materials can be separated and propa-
gated to generate self-similar articial production.
It has been demonstrated that some organic molecules can
control the morphology of biominerals, like calcium phosphate
and calcium carbonate crystals.10
The chiral organic molecules
usually act as templates to control the crystal morphology,
rather than incorporated organic composition to constitute the
chiral hybrid materials.11,12
In the articial design of chiral
nanomaterials, a variety of dispersed chiral superstructures,
such as nano-helixes and nano-tubes, can be generated with the
twisted assembly of chiral molecules, or even nano-sized crys-
talline units.13
However, each nano-helix or nano-tube is con-
structed by independent assembly, rather than a successive
proliferation procedure to pass down the chirality and nal
formation of the homochiral complex. As a result, the archi-
tecture of a homochiral material complex is rarely achieved.14
Herein, by employing a racemic mixture of a chiral amphiphile
(bis-(2-ethylhexyl) sulfosuccinate sodium salt, AOT) and bovine
serum albumin (BSA) in supersaturated calcium phosphate
solution, two kinds of chiral organic–inorganic hybrid nano-
helixes (L- and R-enantiomers) can spontaneously form and
each kind of chiral helix eventually proliferates into a larger
homochiral helix complex. We feel that such an experimental
phenomenon may be relevant to the proliferation of chiral
materials.
2 Experimental section
2.1 Materials
Triply distilled CO2-free water was used in the experiment.
Ca(NO3)2 and (NH4)2HPO4 were of analytical grade and their
solutions were ltered twice through 0.22 mm Millipore lms
prior to use. BSA (Albumin Bovine fraction V, BR, purity  98%)
and AOT (Aldrich, racemic mixture) were directly used without
further purication.
2.2 Preparation of the homochiral nano-helix complex
The temperature during all the synthesis processes was main-
tained at 30 Æ 1 
C. Briey, a 100 ml aqueous solution con-
taining 1 mM AOT and 1 mg mlÀ1
BSA was prepared. The
solution pH was adjusted to 10.0 Æ 0.5 by 3 M ammonia solu-
tion. 50 ml Ca(NO3)2 solution (5 mM, pH ¼ 10.0 Æ 0.5) was
added. Then 50 ml (NH4)2HPO4 solution (3 mM, pH ¼ 10.0 Æ
0.5) was added dropwise at a rate of 1.5 ml minÀ1
. The solution
was gently stirred for 10 h and the formed solids were
collected by centrifugation at 3600 rpm. All the solid samples
were washed by water three times and were vacuum-dried at
35 Æ 1 
C. Freshly prepared samples were dispersed in ethanol
($0.5 mg mlÀ1
) and collected on carbon-coated copper grids for
TEM examinations. In the seed growth experiment, 1/20 percent
of the obtained product underwent intense ultrasonic treat-
ment (KUDOS, 35 kHz, 20 min) and the helix clusters or
networks were collapsed into dispersed helixes. Then the
dispersed helixes were added as seeds into freshly prepared
reaction solutions and the reaction solutions were collected by
centrifugation and observed with Transmission Electron
Microscopy (TEM). For ultrathin sectioned TEM examination,
dried samples were embedded in 0.5 ml epoxy. The mixture was
solidied at 80 
C for 12 h and then carefully microtomed by a
Reichert-Jung Ultracut E ultramicrotome using a diamond
knife.
2.3 Au-labelled BSA absorptions
BSA-Au nanoparticles were synthesized according to the work by
J. Xie et al.15
Briey, 5 ml 10 mM HAuCl4 solution was added into
5 ml 50 mg mlÀ1
BSA solutions and stirred for 5 min. Then,
0.5 ml 1 M NaOH was added and the solution was kept at 37 
C
for 24 h. The product was dialyzed with 1000 ml distilled water
for 24 h. The BSA-Au was used instead of pure BSA in order to
probe the location of BSA on the surface of the helix.
2.4 Examination of calcium concentrations
The concentration of free calcium ions in BSA, AOT, BSA + AOT
solutions were measured by a PCa-1 calcium ion selective
electrode with a saturated calomel electrode as the reference
electrode. The electrode was calibrated according to the
instructions before use.
2.5 Characterizations
Scanning electron microscopy (SEM) was performed by using a
HITACHI S-4800 eld-emission scanning electron microscope
at an acceleration voltage of 5 kV. Fourier-transform infrared
spectroscopy (FT-IR, Nicolet Nexus 670) was used to determine
the composition of the products. Thermogravimetric analysis
(TGA) was carried out by a TA Instrument SDT Q600. The
experiments were measured over a temperature range of 22–
800 
C at a rate of 10 
C minÀ1
under air atmosphere. TEM
observations were performed by a JEM-1200EX at a typical
acceleration voltage of 80 kV. Small angle X-ray diffraction
(SAXRD) and Wide angle X-ray diffraction (WAXRD) were char-
acterized by a Rigaku D/max-2550pc with monochromatized Cu
Ka radiation and the scanning step was 0.02
. Solid state
nuclear magnetic resonance (ssNMR) was kindly performed by
Prof. Jarry Chan's group at the National Taiwan University on a
Bruker DSX300 NMR spectrometer.
3 Results and discussion
3.1 Structure and composition of the nano-helix
In our biomimetic case, AOT and BSA were adopted as the
models for biological amphiphilic and proteins, respectively.
AOT is of asymmetric double-chain amphiphile (Fig. S1†). It can
This journal is ª The Royal Society of Chemistry 2013 Nanoscale, 2013, 5, 3006–3012 | 3007
Paper Nanoscale
Publishedon06February2013.DownloadedbyUniversityofCalifornia-SanFranciscoon24/01/201605:50:31.
View Article Online
assemble into various mesomorphous phases, which have been
widely used in biomimetic crystallization.16
BSA is one of the
common proteins in biomineralization studies.17
The syner-
gistic effect of AOT and BSA on calcium phosphate minerali-
zation gave rise to the formation of nano-helix (Fig. 2 and S2†).
In the control experiments, the use of AOT and BSA alone only
generated the calcium phosphate nanorods and nanospheres,
respectively (Fig. S3†). Clearly, the helix formation was attrib-
uted to the coexistence of AOT, BSA and calcium phosphate.
It followed that the individual helixes could further develop
into micron-sized aggregated clusters (Fig. 2a and b). In an
individual cluster, the nano helixes extended radially outward
from a dense core, indicating the successive proliferation
procedure. Furthermore, some clusters connected with each
other to form a larger network (Fig. 2 and S4†). As the basic
building blocks of the clusters, the helixes were chiral and they
had two kinds of spiral enantiomers, L- and R-forms. Although
the overall amounts of the L- and R-helixes in the reaction
system were equal (Fig. S5†), only one helix enantiomer could be
identied within a cluster or connected network (Fig. 2a–c, see
more in ESI†). This suggested that the spontaneous chiral
recognition and chiral separation occurred during the cluster
and network generation. Concerning the composition and
structure of the helixes, they were constituted by organic and
mineral phases, which accounted for 20.5 wt% and 58.8 wt%,
respectively (the rest 20.7% was attributed to absorbed and
crystal water, Fig. S6†). FT-IR (Fig. 2d) and energy-dispersive
X-ray spectroscopy (EDS, Fig. S7†) revealed that the main
components in the helixes were AOT and a calcium phosphate
phase. X-ray diffraction (Fig. 2e) showed that the mineral phase
was close to brushite. Moreover, the mineral phase in the helix
was conrmed by Multiple Pulse Sequence Nuclear Magnetic
Resonance spectroscopy (CRAMPS-NMR) and a Heteronuclear
Correlation (HETCOR) spectrum between the 31
P and 1
H nuclei
31
P{1
H} combined rotation, indicating that the phosphate
groups were protonated (HPO4
2À
) in the calcium mineral
(Fig. 2f, S8 and Table S1†). The NMR data indicated the absence
of PO4
3À
in the complex. As a result, the calcium phosphate
species containing PO4
3À
groups such as hydroxyapatite, octa-
calcium phosphate and tri-calcium phosphate could be
excluded in the phase analysis. Additionally, Ca(H2PO4)2 could
not be precipitated under our experimental conditions due to
its high solubility. Both examinations shows that the signals of
the helix were close to those of brushite. Therefore, the brush-
ite-like mineral was considered as the primary inorganic
component in the helixes.
The internal structure of the nano-helixes could be consid-
ered as the alternative and spiral stacking of thin calcium
phosphate phase and AOT bilayers. The cross-section images of
the nano-helix showed that the thickness of the wall of the
nano-helix was about 2.1–2.2 nm (Fig. 3a–c). Furthermore,
SAXRD and WAXRD results also showed the alternative lamellar
superstructure in the helixes with a constant interspacing
distance (d ¼ 3.34 nm). The lamellar structure was also
demonstrated by TEM (Fig. 2e): the dark lines (1.7 nm) and light
lines (1.6 nm) correspond to the inorganic calcium phosphate
and organic AOT ultrathin layers, respectively. The thickness of
each organic–inorganic hybrid unit was about 3.3 Æ 0.2 nm,
which is in agreement with the d value calculated from the
SAXRD data. In the spiral helix, there existed a pitch angle of
about 43
between the strip edge and the long axis. It was noted
that the AOT molecules preferred to assemble into a bilayer
structure. Concisely, the organic bilayer could have a thickness
of about 1.6 nm if the molecules tilted by 43
. There were two
mirror forms for both the helix pitch angle and the AOT tilt
angle, +43
or À43
, as the denitions (Fig. 3d and e). The
mirror packing of AOT corresponded to the formation of R- and
L-enantiomers of the helixes. Apart from AOT, a small amount
of BSA was detected in the helix by FT-IR and 13
C{1
H} NMR
(Fig. S9†). Using nano Au particle labelled BSA as the imaging
agent, we found that the protein did not incorporate into the
hybrid inner structure, but absorbed onto the helix wall
surfaces (Fig. S10†), which might be due to its relatively large
dimension.19
We suggested that BSA served as a surface or
Fig. 2 Characterizations of nanohelixes. (a and b) Homochiral clusters consisting
of R-helixes and L-helixes, respectively. (c) A homochiral helix network; circles
indicate the cluster centres; inset is a magnification of the rectangular region. (d)
FT-IR of the helix and pure AOT. The typical and undisturbed peaks of AOT
(1750 cmÀ1
), BSA (1540 cmÀ1
, amino) and phosphate ions could be noted. The
peak located at 2342 cmÀ1
was generally attributed to CO2 from the air during the
FT-IR determination. (e) SAXRD and WAXRD patterns of the helixes. d ¼ 3.34 nm
and d ¼ 1.65 nm represent the first and the second diffractions of the lamellar
structure in the helixes, respectively. WAXRD also showed that the mineral phase
was similar to brushite. The XRD peaks of 11.3
and 31.0
were close to those of
brushite (020) and (121), respectively. The case of the small left-shift of charac-
teristic peaks could be found in small nanocrystals.18
The inset TEM image shows
each organic (light line)–inorganic (dark line) unit in the helix. (f) 31
P{1
H}HETCOR
spectra between the 31
P and 1
H nuclei measured in the helixes. The spectra was
acquired at a spinning frequency of 10 kHz and the contact time was set to 2.5 ms.
A total of 64 transients with an increment of 100 ms was accumulated.
3008 | Nanoscale, 2013, 5, 3006–3012 This journal is ª The Royal Society of Chemistry 2013
Nanoscale Paper
Publishedon06February2013.DownloadedbyUniversityofCalifornia-SanFranciscoon24/01/201605:50:31.
View Article Online
structure stabilizer for the hybrid spiral strips. An experimental
fact was that no chiral product formed by using BSA alone.
During biomineralization, some biomacromolecules can adopt
an extended conformation when they interact with the inor-
ganic phase surface.20
Therefore, there was a possibility that the
BSA molecules might incorporate into the helix using their
extended forms. However, we could conrm that the AOT
molecules with highly charged sulphuric groups, rather than
BSA, were the primary organic composition in the nano-helix.
For example, the FT-IR study showed a very weak amino peak
(Fig. 2d) in the composites, implying the ignorable contents of
BSA in comparison with the strong AOT bands. Therefore, it was
suggested that the hybrid helixes were formed by the ordered
assembly of the calcium phosphate mineralized layer and the
AOT bilayer, and were stabilized by BSA absorption on the
hybrid surfaces. In this architecture, the assembly behaviour of
chiral AOT molecules in the hybrid helixes determined the
material’s chirality.
3.2 Proliferation of the nano-helixes
Originally, the homochiral clusters and networks evolved from a
single nano-helix (mother matrix). The preformed chiral nano-
helix spontaneously passed down the structure information
(chirality) from one generation to the next and then generated
the homochiral complexes (Fig. 4). Firstly, tiny hybrid buds
sprouted from the surface of the matrix (Fig. 4a). Both organic
and inorganic parts of the buds directly integrated with the
corresponding parts of mother matrix. This could be considered
as a kind of matrix outgrowth. Secondly, these hybrid buds grew
longer and twisted into the helical ribbons. At this stage, the
organic and inorganic parts at the growing front of hybrid buds
did not integrate with mother matrix any more. At this time,
there should be a choice in the twist direction (L- or R-).
Nevertheless, we noted that the newly formed helical ribbons
replicated precisely the twist direction (chirality) of the mother
matrix. This meant that the mother matrix induced the later
AOT molecules to assemble into a coherent packing direction,
even the AOT bilayers at the budding region and growth front
were separated by calcium phosphate layers. Thus, the original
structure was inherited through the budding and proliferation
process (Fig. 4b). Thirdly, the homochiral proliferation process
of the helixes continued by generating more “daughters” and
“grand-daughters” based upon the matrix. Due to the space
limitation, the newly formed helixes tended to stretch outward,
which generated radial homochiral clusters (Fig. 4c). Finally, a
few of the helixes at the cluster edge acted as “bridges” to
provide additional growing sites for new buds and initiated
another proliferation process (Fig. 4e). This new proliferation
Fig. 3 (a and b) Cross-section images of nano-helixes under TEM. (c) Schematic
structure of the nano helixes (dark grey: inorganic phase; light grey: organic
phase). (d and e) TEM and schemes of the R- and L-helix. The width of the AOT
bilayer is 1.6 nm from TEM observation. As AOT molecules have a length of 1.1
nm, AOT molecules in a bilayer should arrange with a tilt angle of about 43
.
Note: the AOT molecules in the same bilayer are simply treated as direct contact
and this small variation of tilt angle doesn't affect our qualitative analysis.
Fig. 4 TEM images of the evolution from a single helix to a homochiral cluster
and then a homochiral network (community). (a) A sprouting bud from R-helix
matrix for the new “daughter” helix generation. (b) Growth and twist of the
“daughter” helix, which duplicated the chiral feature to be R-form; insets show
the details of the growth front on the matrix. (c) More buds formed and they
replicated the structure of matrix precisely. (d) Rudiment of the homochiral helix
cluster; insets: magnification of the branching sites. (e) Homochiral helix cluster
(R-form); arrows indicate the proliferation directions of the cluster; inset shows
the new buds formed at an extended helix. (f) Homochiral helix networks (R-
form); arrows show the proliferation directions.
This journal is ª The Royal Society of Chemistry 2013 Nanoscale, 2013, 5, 3006–3012 | 3009
Paper Nanoscale
Publishedon06February2013.DownloadedbyUniversityofCalifornia-SanFranciscoon24/01/201605:50:31.
View Article Online
could happen at multiple directions (Fig. 4f). Through the self-
repeating processes, a single nano-sized helix eventually evolved
into a large homochiral complex (network) at the micrometer
scale (Fig. 4f). In each network, the chirality of the newly born
helixes was precisely “inherited” from the original mother
matrix from generation to generation, which can be considered
as a spontaneous process of material-based self-proliferation.
We note that it is impossible for the dispersed helixes aer
intense ultrasonic treatment to aggregate into homochiral
clusters again. However, aer the dispersed helixes are re-
dispersed into the freshly prepared reaction solutions as seeds,
the time to induce the formation of helix clusters can be rela-
tively reduced according to the seed amount, indicating that the
mother helix acts as the seed to induce the proliferation of new
helixes to form helix clusters (Fig. S11†). As a result, in each
network, the chirality of the newly born helixes was precisely
“inherited” from the original mother matrix from generation to
generation, which can be considered as a spontaneous process
of material-based self-proliferation.
3.3 Model for the homochiral material
Analogous to the shell formation, the co-assembly of organic and
inorganic phases is restricted at the local domains of the growth
front (Fig. 4a and b). AOT molecules are more able of binding
calcium ions than BSA (Table S2†) and these amphiphilic
molecules greatly modulate the growth of calcium phosphate
species.21
In this case, the helix formation and replication are
controlled dominantly by the assembly behaviours of AOT at the
growth front. The AOT molecules can tightly absorb onto the
surface of calcium phosphate species with a strong binding
effect between calcium ions and sulphuric groups, which facili-
tates the assembly of AOT bilayers.22
Unlike the free AOT mole-
cules in aqueous solution, the relatively rigid CaP crystal, rather
than the mobile water layer, can imobilize the adjacent AOT
bilayers. Thus, the hybrid structure could be ‘solidied’ and
stabilized with a decreasing disordered uctuation of AOT
molecules comparing the free state, which then induces the next
layer mineralization.23
This alternative and cooperated deposi-
tion of the AOT bilayer and calcium phosphate phase layer
gradually constitutes a thin AOT–calcium phosphate hybrid
strip, which is similar to the associated assembly of lipids and
inorganic phase reported by Seddon, et.al.13b
In our work, the chiral AOT molecules are responsible for the
twist of the hybrid strip to form the L- and R-nano-helixes.
Although the BSA used here is constituted with chiral L-amino
acids, the chirality of nano-helixes are unlikely to be controlled
by BSA. Due to its large size, it is difficult to incorporate into the
ordered structure with ultra-small units of 1.7 nm (calcium
phosphate layer) and 1.6 nm (organic layer), while the twisted
arrangement of these units forms the chirality at the nanoscale.
In addition, the equal number of L- and R-nano-helixes also
indicates that the BSA with single chiral units (L-amino acids)
has little contribution to the chirality of the nano-helixes. Many
works have been reported that strips constituted with chiral
molecules tend to twist into nano-helixes to reduce the elastic
energy.24
Similarly, the chirality of the nano-helixes in our
system is determined by the assembly behaviour of chiral AOT
molecules.
However, the racemic mixture generally dilutes the chiral
interaction between the chiral molecules, so that the chiral
superstructure might fail to form.25
Nevertheless, some studies
have shown that both R- and L-enantiomers can emerge in
racemic systems if an energy favoured chiral phase separation
occurs, especially for the lipids with chiral headgroups and
inexible double chains structures.26
Interestingly, AOT owns a
similar structure and phase behaviour to these lipids.27
More-
over, chiral molecules can also undergo a phase separation
when they are restricted at interfaces.28
Therefore, in our
system, it follows that a spontaneous chiral phase separation of
amphiphilic AOT may occur on the calcium phosphate mineral
substrate, resulting in the bilayers with exactly the same
molecular packing behaviour.
Due to the complicated structure, the conformation infor-
mation (chirality) of AOT in each nano-sized helix is difficult to
identify. Besides, methods of the synthesis or separation of AOT
diastereoisomers is rarely reported.29
Based upon the mirror
arrangement of AOT in L- and R-helixes, we divide the AOT
molecules into two types with different tilt directions of +43
or À43
. Aer this simplication, only the tilt angle needs to be
taken into account in the qualitative analysis of the energy
during the formation process. The AOT molecules in the bila-
yers have two different tilt angles, +43
or À43
, which can be
considered as the enantiomers to induce R- and L-chiral helix
formations, respectively (Fig. 3d and e). The favoured tilt angle
should maintain the same value during the alternative dispo-
sition. Thus, the energy favoured recognition is a key to main-
tain the molecular assembly according to the chiral breaking
model supposed by Selinger et al.30
The model suggests that the
elastic energy of the strip can be reduced by a chiral separation
even under racemic conditions. An order parameter, j, is
introduced, which is treated as the local net amount of right-
handed minus le-handed molecular packing here. The elastic
free energy, F, of the thin chiral bilayer strip can be written as
eqn (1)
F ¼
ð
dS

1
2
k

1
r
2
þ
1
2
k0

1
r
2
cos2
f À lHPj

1
r

sin f cos f
þ
1
2
KðVjÞ2
þ
1
2
tj2
þ
1
4
uj4

þ Eedge (1)
where, S is the area, the rst term is the standard Helfrich
bending energy of the hybrid membrane and the coefficient k is
the isotropic rigidity. In the right side of eqn (1), the second
term represents the anisotropy of the rigidity and the coefficient
k0
is the anisotropic term and f is the title angle of the chiral
molecules (Fig. 5a); the third term is a chiral term that favours
twisting in a tilt angle f; the coefficient lHP, is the chirality
parameter, which exists only in chiral membranes and depends
on the chiral order. The sign of lHP can be changed when the
membrane transforms into its mirror image. lHPj increases
with the greater chiral phase separation degree of j. The last
three terms in the bracket are the Ginzburg–Landau expres-
sions in powers of j, which represent the free energy change
3010 | Nanoscale, 2013, 5, 3006–3012 This journal is ª The Royal Society of Chemistry 2013
Nanoscale Paper
Publishedon06February2013.DownloadedbyUniversityofCalifornia-SanFranciscoon24/01/201605:50:31.
View Article Online
during the ordering transition. The values of K and u are
temperature independent constants while the coefficient of t
relates to temperature and t  0 for chiral phase separation.30,31
Because AOT molecules in the helix have the xed tilt angle of
43
, the domain wall energy on the edge is a constant.
When simultaneously minimizing the free energy over tilt
angle and radius, r, in eqn (1), the following is obtained,
f0 ¼ arctan

k þ k0
k
1=
4

(2)
r0 ¼
k
1=
4
ðk þ k0
Þ
1=
4
h
k
1=
2
þ ðk þ k0
Þ
1=
2
i
lHPj
(3)
in which, (k + k0
)/k represents the energy cost for the ratio of the
bend parallel to the tilt direction to bend perpendicular to
the tilt direction. In our system, its value is about 0.76 because
fAOT ¼ 43
, which indicates that the hybrid strip favours
twisting parallel to the AOT tilt direction. Besides, the radius of
the nano-helix equals 1.33k/lHPj. Usually, the lipid amphi-
philes form helical tubes or a helix with a larger diameter of
hundreds nanometres or even a few micrometers. Since the
organic–inorganic hybrid structure exists in our helixes, it is
reasonable that the rigidity coefficient k should be greater than
single component chiral lipid membranes. As a result, lHPj
must be signicantly greater to produce such a slender helix
with a very small radius ($10 to 25 nm).
We note that the favoured energy barrier DF plays an
essential role to control the twist direction and chiral prolifer-
ation of nano-helixes (eqn (1)). Here, the qualitative description
energy barrier DF between the racemic state (j ¼ 0) and sepa-
ration state (jmÆ) can be described by eqn (4),30
DF ¼
ð
dS

lHPjmÆ

1
r

sin f cos f þ
1
2
KðVjmÆÞ2
þ
1
2
tjmÆ
2
þ
1
4
ujmÆ
4

(4)
As the radius of the nano-helix is r f 1/lHPj, the relatively
small radius of the helix (10–25 nm) implies that lHPj is of great
value in our case, which facilitates the chiral separation. The
equation shows that the free energy of the strip has two local
minima representing the two types of energy favoured AOT
packing with the mirror symmetry (Fig. 5b) if chiral phase
separation occurs. Without chiral phase separation, the strip
cannot twist into a helix because the radius becomes innite
when j ¼ 0 (r / N).
Fig. 5b shows the constant arrangement of AOT molecules
(constant tilt angle of +43
or À43
) within a strip is energeti-
cally preferred due to an energy barrier. First, the energy barrier
DF can promote the formation of energy favoured and stable
nano-helixes, rather than unstable hybrid strips. If the different
arrangements of AOT molecules (L- and R-) coexist in the same
strip, the elastic energy increases so that the resulting strip
becomes unstable (Fig. 5b, middle). Therefore, the elastic
energy cannot be reduced and generate twisted nano-helixes. By
contrast, the same arrangements of AOT molecules (L- or R-) can
successfully reduce the unfavoured elastic energy and twist to
form L- or R-nano-helixes, respectively (Fig. 5b, le and right).
Second, the energy barrier of DF is also responsible for the
homochiral proliferation. In our system, the preformed helix
matrix has an inductive effect on the sequent proliferation
because the emerging organic and inorganic parts in the new
buds directly extend from their mother matrix. Thus, new buds
share the same AOT packing form with the mother matrix. The
same AOT packing can be replicated under the guidance of the
mother matrix due to the favoured energy reduction, which
means that L- to L- or R- to R-proliferation is a preferential way.
Subsequently, the buds grow following the determined AOT
packing to form a new chiral helix with the same chirality. For
example, the new buds generated from the R-nano-helixes in
Fig. 4 faithfully adopt the R-twist direction and keep the selected
form during the growth process. The mutated proliferation of
L- to R- or R- to L- also require extra energy to overcome DF in
comparison with the matched L- to L- or R- to R-. Accordingly,
the chiral structure proliferation always initiates at the pre-
formed helixes and amplies the chiral structure from the
mother matrix to subsequent generations. Finally, large
homochiral complexes (helix clusters and networks) can be
generated under the guidance of the energy controlled recog-
nition of AOT packing.
4 Conclusions
This study reveals that the homochiral complex of the organic–
inorganic hybrid helix can form via a self-proliferation process.
The energy controlled chiral recognitions and separations of
asymmetric chiral AOT molecules are essential in both helix
formation and homochiral proliferation. The nding is of
importance to approach homochiral biomimetic materials in
the laboratory. We expect this strategy of bio-inspired chiral
structure proliferation can be developed into a convenient
pathway for the articial synthesis of self-similar functional
materials.
Acknowledgements
We thank Prof. Jerry Chen for the ssNMR studies, Dr Jinhui Tao,
Dr Haihua Pan and Yuan Su for discussions, Hua Wang, Jieru
Fig. 5 (a) The geometry of AOT molecules in the helix discussed in eqn (1) for helix
formation. (b) Two local minima of the elastic free energy (F) with symmetry
packing (jm+ or jmÀ) lead to an energy barrier of DF, which ensures the oriented
packing vector of AOT bilayers to produce chiral helix and homochiral proliferation.
This journal is ª The Royal Society of Chemistry 2013 Nanoscale, 2013, 5, 3006–3012 | 3011
Paper Nanoscale
Publishedon06February2013.DownloadedbyUniversityofCalifornia-SanFranciscoon24/01/201605:50:31.
View Article Online
5. all published paper
5. all published paper
5. all published paper
5. all published paper
5. all published paper
5. all published paper
5. all published paper
5. all published paper
5. all published paper
5. all published paper
5. all published paper
5. all published paper
5. all published paper
5. all published paper
5. all published paper
5. all published paper
5. all published paper
5. all published paper
5. all published paper
5. all published paper
5. all published paper
5. all published paper
5. all published paper
5. all published paper
5. all published paper
5. all published paper
5. all published paper
5. all published paper
5. all published paper
5. all published paper
5. all published paper
5. all published paper

More Related Content

What's hot

synthesis Of metal organic frame work ( MOFs)
synthesis Of metal organic frame work ( MOFs)synthesis Of metal organic frame work ( MOFs)
synthesis Of metal organic frame work ( MOFs)University Of Wah
 
ACETYLATION OF BENZYLIC ALCOHOLS OVER BiFeO3 (BFO), Bi0.86Sm0.07Eu0.07FeO3 (B...
ACETYLATION OF BENZYLIC ALCOHOLS OVER BiFeO3 (BFO), Bi0.86Sm0.07Eu0.07FeO3 (B...ACETYLATION OF BENZYLIC ALCOHOLS OVER BiFeO3 (BFO), Bi0.86Sm0.07Eu0.07FeO3 (B...
ACETYLATION OF BENZYLIC ALCOHOLS OVER BiFeO3 (BFO), Bi0.86Sm0.07Eu0.07FeO3 (B...EDITOR IJCRCPS
 
Highly Tunable Molecular Sieving and Adsorption Properties of Mixed-Linker Ze...
Highly Tunable Molecular Sieving and Adsorption Properties of Mixed-Linker Ze...Highly Tunable Molecular Sieving and Adsorption Properties of Mixed-Linker Ze...
Highly Tunable Molecular Sieving and Adsorption Properties of Mixed-Linker Ze...Krishna Chandran Jayachandrababu
 
Sensing of volatile organic compounds by MOFs
Sensing of volatile organic compounds by MOFsSensing of volatile organic compounds by MOFs
Sensing of volatile organic compounds by MOFsMohammadRad12
 
Omar Yaghi at BASF Science Symposium 2015
Omar Yaghi at BASF Science Symposium 2015Omar Yaghi at BASF Science Symposium 2015
Omar Yaghi at BASF Science Symposium 2015BASF
 
Nano research 2014 qian liu
Nano research 2014 qian liuNano research 2014 qian liu
Nano research 2014 qian liuQian Liu, phD
 
MoFs (METAL-ORGANIC FRAMEWORKS)
MoFs (METAL-ORGANIC FRAMEWORKS)MoFs (METAL-ORGANIC FRAMEWORKS)
MoFs (METAL-ORGANIC FRAMEWORKS)mahsaehsani
 
1 s2.0-s0169131717302016-main
1 s2.0-s0169131717302016-main1 s2.0-s0169131717302016-main
1 s2.0-s0169131717302016-maingiuliani1
 
Synthesis of HTTP-MOFs on HOPG Poster
Synthesis of HTTP-MOFs on HOPG PosterSynthesis of HTTP-MOFs on HOPG Poster
Synthesis of HTTP-MOFs on HOPG PosterAlice Hsu
 
CAPTURING NEON IN A METAL-ORGANIC FRAMEWORK
CAPTURING NEON IN A METAL-ORGANIC FRAMEWORKCAPTURING NEON IN A METAL-ORGANIC FRAMEWORK
CAPTURING NEON IN A METAL-ORGANIC FRAMEWORKCCDC
 
Self assembled monolayers
Self assembled monolayersSelf assembled monolayers
Self assembled monolayersSumit Kumar
 
Surface Modification of Nanoparticles for Biomedical Applications
Surface Modification of Nanoparticles for Biomedical ApplicationsSurface Modification of Nanoparticles for Biomedical Applications
Surface Modification of Nanoparticles for Biomedical ApplicationsReset_co
 
Hydrothermal synthesis and characterization of one
Hydrothermal synthesis and characterization of oneHydrothermal synthesis and characterization of one
Hydrothermal synthesis and characterization of oneAlexander Decker
 
Covalant Organic Frameworks
Covalant Organic FrameworksCovalant Organic Frameworks
Covalant Organic FrameworksMuhammad Ehsan
 
Applications of Metal Organic Frameworks(MOF) in water treatment
Applications of Metal Organic Frameworks(MOF) in water treatmentApplications of Metal Organic Frameworks(MOF) in water treatment
Applications of Metal Organic Frameworks(MOF) in water treatmentSainabijanzadeh
 
Ijms 15-07639 solucion-celulascancer
Ijms 15-07639 solucion-celulascancerIjms 15-07639 solucion-celulascancer
Ijms 15-07639 solucion-celulascancerakire_aicrag22
 
Effect of Solvents on Size and Morphologies Of sno Nanoparticles via Chemical...
Effect of Solvents on Size and Morphologies Of sno Nanoparticles via Chemical...Effect of Solvents on Size and Morphologies Of sno Nanoparticles via Chemical...
Effect of Solvents on Size and Morphologies Of sno Nanoparticles via Chemical...Editor IJCATR
 

What's hot (20)

synthesis Of metal organic frame work ( MOFs)
synthesis Of metal organic frame work ( MOFs)synthesis Of metal organic frame work ( MOFs)
synthesis Of metal organic frame work ( MOFs)
 
ACETYLATION OF BENZYLIC ALCOHOLS OVER BiFeO3 (BFO), Bi0.86Sm0.07Eu0.07FeO3 (B...
ACETYLATION OF BENZYLIC ALCOHOLS OVER BiFeO3 (BFO), Bi0.86Sm0.07Eu0.07FeO3 (B...ACETYLATION OF BENZYLIC ALCOHOLS OVER BiFeO3 (BFO), Bi0.86Sm0.07Eu0.07FeO3 (B...
ACETYLATION OF BENZYLIC ALCOHOLS OVER BiFeO3 (BFO), Bi0.86Sm0.07Eu0.07FeO3 (B...
 
Highly Tunable Molecular Sieving and Adsorption Properties of Mixed-Linker Ze...
Highly Tunable Molecular Sieving and Adsorption Properties of Mixed-Linker Ze...Highly Tunable Molecular Sieving and Adsorption Properties of Mixed-Linker Ze...
Highly Tunable Molecular Sieving and Adsorption Properties of Mixed-Linker Ze...
 
Sensing of volatile organic compounds by MOFs
Sensing of volatile organic compounds by MOFsSensing of volatile organic compounds by MOFs
Sensing of volatile organic compounds by MOFs
 
Omar Yaghi at BASF Science Symposium 2015
Omar Yaghi at BASF Science Symposium 2015Omar Yaghi at BASF Science Symposium 2015
Omar Yaghi at BASF Science Symposium 2015
 
Nano research 2014 qian liu
Nano research 2014 qian liuNano research 2014 qian liu
Nano research 2014 qian liu
 
MoFs (METAL-ORGANIC FRAMEWORKS)
MoFs (METAL-ORGANIC FRAMEWORKS)MoFs (METAL-ORGANIC FRAMEWORKS)
MoFs (METAL-ORGANIC FRAMEWORKS)
 
1 s2.0-s0169131717302016-main
1 s2.0-s0169131717302016-main1 s2.0-s0169131717302016-main
1 s2.0-s0169131717302016-main
 
Synthesis of HTTP-MOFs on HOPG Poster
Synthesis of HTTP-MOFs on HOPG PosterSynthesis of HTTP-MOFs on HOPG Poster
Synthesis of HTTP-MOFs on HOPG Poster
 
CAPTURING NEON IN A METAL-ORGANIC FRAMEWORK
CAPTURING NEON IN A METAL-ORGANIC FRAMEWORKCAPTURING NEON IN A METAL-ORGANIC FRAMEWORK
CAPTURING NEON IN A METAL-ORGANIC FRAMEWORK
 
Joe 02 2005_97_100
Joe 02 2005_97_100Joe 02 2005_97_100
Joe 02 2005_97_100
 
SIRFinalPaper1
SIRFinalPaper1SIRFinalPaper1
SIRFinalPaper1
 
Self assembled monolayers
Self assembled monolayersSelf assembled monolayers
Self assembled monolayers
 
Surface Modification of Nanoparticles for Biomedical Applications
Surface Modification of Nanoparticles for Biomedical ApplicationsSurface Modification of Nanoparticles for Biomedical Applications
Surface Modification of Nanoparticles for Biomedical Applications
 
Hydrothermal synthesis and characterization of one
Hydrothermal synthesis and characterization of oneHydrothermal synthesis and characterization of one
Hydrothermal synthesis and characterization of one
 
Covalant Organic Frameworks
Covalant Organic FrameworksCovalant Organic Frameworks
Covalant Organic Frameworks
 
Applications of Metal Organic Frameworks(MOF) in water treatment
Applications of Metal Organic Frameworks(MOF) in water treatmentApplications of Metal Organic Frameworks(MOF) in water treatment
Applications of Metal Organic Frameworks(MOF) in water treatment
 
Ijms 15-07639 solucion-celulascancer
Ijms 15-07639 solucion-celulascancerIjms 15-07639 solucion-celulascancer
Ijms 15-07639 solucion-celulascancer
 
Effect of Solvents on Size and Morphologies Of sno Nanoparticles via Chemical...
Effect of Solvents on Size and Morphologies Of sno Nanoparticles via Chemical...Effect of Solvents on Size and Morphologies Of sno Nanoparticles via Chemical...
Effect of Solvents on Size and Morphologies Of sno Nanoparticles via Chemical...
 
Metal organic framework
Metal organic frameworkMetal organic framework
Metal organic framework
 

Viewers also liked

Rx swift,Repro framewolrk
Rx swift,Repro framewolrkRx swift,Repro framewolrk
Rx swift,Repro framewolrkDaisuke Nagata
 
Gender inequality in media
Gender inequality in mediaGender inequality in media
Gender inequality in mediaaluxbaiya
 
Carlosjose ramirez hardware
Carlosjose ramirez hardwareCarlosjose ramirez hardware
Carlosjose ramirez hardwarecarlos ramirez
 
best web design company noida%9999623343%web design company in delhi ncr
best web design company noida%9999623343%web design company in delhi ncrbest web design company noida%9999623343%web design company in delhi ncr
best web design company noida%9999623343%web design company in delhi ncrbittarget1
 
Yuliana carrero hardware.ppt
Yuliana carrero hardware.pptYuliana carrero hardware.ppt
Yuliana carrero hardware.pptyuliana carrero
 
Tecnologia formacao profissional_unidade1
Tecnologia formacao profissional_unidade1Tecnologia formacao profissional_unidade1
Tecnologia formacao profissional_unidade1Ramon Lucena
 
SHERRI'S RESUME REVISED##
SHERRI'S RESUME REVISED##SHERRI'S RESUME REVISED##
SHERRI'S RESUME REVISED##Sherri Allen
 
Bachillerato general EMILIANO ZAPATAcuetzalan
Bachillerato general  EMILIANO ZAPATAcuetzalanBachillerato general  EMILIANO ZAPATAcuetzalan
Bachillerato general EMILIANO ZAPATAcuetzalanRamon99alonso29
 
Goods and service act (1)
Goods and service act (1)Goods and service act (1)
Goods and service act (1)Joy Waghela
 
Bittarget digital marketing-campaign in noida
Bittarget digital marketing-campaign in noidaBittarget digital marketing-campaign in noida
Bittarget digital marketing-campaign in noidabittarget1
 
CV T Mooi June 2016
CV T Mooi June 2016CV T Mooi June 2016
CV T Mooi June 2016Trevor Mooi
 

Viewers also liked (20)

La ergonomía en la organización de la empresa
La ergonomía en la organización de la empresaLa ergonomía en la organización de la empresa
La ergonomía en la organización de la empresa
 
Presentación1caledonia (1)
Presentación1caledonia (1)Presentación1caledonia (1)
Presentación1caledonia (1)
 
Rx swift,Repro framewolrk
Rx swift,Repro framewolrkRx swift,Repro framewolrk
Rx swift,Repro framewolrk
 
Presentacion power point lalo
Presentacion power point laloPresentacion power point lalo
Presentacion power point lalo
 
Gender inequality in media
Gender inequality in mediaGender inequality in media
Gender inequality in media
 
Carlosjose ramirez hardware
Carlosjose ramirez hardwareCarlosjose ramirez hardware
Carlosjose ramirez hardware
 
best web design company noida%9999623343%web design company in delhi ncr
best web design company noida%9999623343%web design company in delhi ncrbest web design company noida%9999623343%web design company in delhi ncr
best web design company noida%9999623343%web design company in delhi ncr
 
Yuliana carrero hardware.ppt
Yuliana carrero hardware.pptYuliana carrero hardware.ppt
Yuliana carrero hardware.ppt
 
Caso chantaje
Caso chantajeCaso chantaje
Caso chantaje
 
Tecnologia formacao profissional_unidade1
Tecnologia formacao profissional_unidade1Tecnologia formacao profissional_unidade1
Tecnologia formacao profissional_unidade1
 
RISTORI
RISTORIRISTORI
RISTORI
 
SHERRI'S RESUME REVISED##
SHERRI'S RESUME REVISED##SHERRI'S RESUME REVISED##
SHERRI'S RESUME REVISED##
 
Presentación1caledonia (1)
Presentación1caledonia (1)Presentación1caledonia (1)
Presentación1caledonia (1)
 
Tugas sim kurniawan
Tugas sim kurniawanTugas sim kurniawan
Tugas sim kurniawan
 
The Hover: Software for building web and mobile apps
The Hover: Software for building web and mobile appsThe Hover: Software for building web and mobile apps
The Hover: Software for building web and mobile apps
 
Bachillerato general EMILIANO ZAPATAcuetzalan
Bachillerato general  EMILIANO ZAPATAcuetzalanBachillerato general  EMILIANO ZAPATAcuetzalan
Bachillerato general EMILIANO ZAPATAcuetzalan
 
Goods and service act (1)
Goods and service act (1)Goods and service act (1)
Goods and service act (1)
 
CCTyson2016Resume
CCTyson2016ResumeCCTyson2016Resume
CCTyson2016Resume
 
Bittarget digital marketing-campaign in noida
Bittarget digital marketing-campaign in noidaBittarget digital marketing-campaign in noida
Bittarget digital marketing-campaign in noida
 
CV T Mooi June 2016
CV T Mooi June 2016CV T Mooi June 2016
CV T Mooi June 2016
 

Similar to 5. all published paper

Davey l1 macromolec-struc-anlys(1) lec 1
Davey l1 macromolec-struc-anlys(1) lec 1Davey l1 macromolec-struc-anlys(1) lec 1
Davey l1 macromolec-struc-anlys(1) lec 1RANJANI001
 
Structural characterization of TiO2 films grown on LaAlO3 and SrTiO3 substrat...
Structural characterization of TiO2 films grown on LaAlO3 and SrTiO3 substrat...Structural characterization of TiO2 films grown on LaAlO3 and SrTiO3 substrat...
Structural characterization of TiO2 films grown on LaAlO3 and SrTiO3 substrat...Oleg Maksimov
 
Soft Matter Algal Adhesive
Soft  Matter  Algal  AdhesiveSoft  Matter  Algal  Adhesive
Soft Matter Algal Adhesiveuvperson
 
Article2016
Article2016Article2016
Article2016jabraoui
 
Formation SiO2 Mass-Independent Oxygen Isotopic Partitioning During Gas-Phase
 Formation SiO2 Mass-Independent Oxygen Isotopic Partitioning During Gas-Phase Formation SiO2 Mass-Independent Oxygen Isotopic Partitioning During Gas-Phase
Formation SiO2 Mass-Independent Oxygen Isotopic Partitioning During Gas-PhaseCarlos Bella
 
Synthesis, characterization and DC Conductivity Studies of NaCF3SO3 -PEO Dope...
Synthesis, characterization and DC Conductivity Studies of NaCF3SO3 -PEO Dope...Synthesis, characterization and DC Conductivity Studies of NaCF3SO3 -PEO Dope...
Synthesis, characterization and DC Conductivity Studies of NaCF3SO3 -PEO Dope...IJERA Editor
 
Identification of physical origin behind disorder, heterogeneity, and reconst...
Identification of physical origin behind disorder, heterogeneity, and reconst...Identification of physical origin behind disorder, heterogeneity, and reconst...
Identification of physical origin behind disorder, heterogeneity, and reconst...Taame Abraha Berhe
 
Multi-molecular views of a stellar nursery
Multi-molecular views of a stellar nurseryMulti-molecular views of a stellar nursery
Multi-molecular views of a stellar nurserySérgio Sacani
 
Non-parabolicity and band gap renormalisation in Si doped ZnO
Non-parabolicity and band gap renormalisation in Si doped ZnONon-parabolicity and band gap renormalisation in Si doped ZnO
Non-parabolicity and band gap renormalisation in Si doped ZnOUniversity of Liverpool
 
Assemblies of aziridinemethanols
Assemblies of aziridinemethanolsAssemblies of aziridinemethanols
Assemblies of aziridinemethanolsPeter ten Holte
 
B.Zhao paper1-chem.mater-2005 SWNT bone mineralization
B.Zhao paper1-chem.mater-2005 SWNT bone mineralizationB.Zhao paper1-chem.mater-2005 SWNT bone mineralization
B.Zhao paper1-chem.mater-2005 SWNT bone mineralizationbinzhao2004
 
Energy-Dispersive X-Ray Spectroscopy Procedure for Analysing Cellular Element...
Energy-Dispersive X-Ray Spectroscopy Procedure for Analysing Cellular Element...Energy-Dispersive X-Ray Spectroscopy Procedure for Analysing Cellular Element...
Energy-Dispersive X-Ray Spectroscopy Procedure for Analysing Cellular Element...iosrjce
 

Similar to 5. all published paper (20)

Davey l1 macromolec-struc-anlys(1) lec 1
Davey l1 macromolec-struc-anlys(1) lec 1Davey l1 macromolec-struc-anlys(1) lec 1
Davey l1 macromolec-struc-anlys(1) lec 1
 
Aquasomes
AquasomesAquasomes
Aquasomes
 
Aquasomes
AquasomesAquasomes
Aquasomes
 
Structural characterization of TiO2 films grown on LaAlO3 and SrTiO3 substrat...
Structural characterization of TiO2 films grown on LaAlO3 and SrTiO3 substrat...Structural characterization of TiO2 films grown on LaAlO3 and SrTiO3 substrat...
Structural characterization of TiO2 films grown on LaAlO3 and SrTiO3 substrat...
 
Soft Matter Algal Adhesive
Soft  Matter  Algal  AdhesiveSoft  Matter  Algal  Adhesive
Soft Matter Algal Adhesive
 
Article2016
Article2016Article2016
Article2016
 
Formation SiO2 Mass-Independent Oxygen Isotopic Partitioning During Gas-Phase
 Formation SiO2 Mass-Independent Oxygen Isotopic Partitioning During Gas-Phase Formation SiO2 Mass-Independent Oxygen Isotopic Partitioning During Gas-Phase
Formation SiO2 Mass-Independent Oxygen Isotopic Partitioning During Gas-Phase
 
Walton-2015-Nature
Walton-2015-NatureWalton-2015-Nature
Walton-2015-Nature
 
Synthesis, characterization and DC Conductivity Studies of NaCF3SO3 -PEO Dope...
Synthesis, characterization and DC Conductivity Studies of NaCF3SO3 -PEO Dope...Synthesis, characterization and DC Conductivity Studies of NaCF3SO3 -PEO Dope...
Synthesis, characterization and DC Conductivity Studies of NaCF3SO3 -PEO Dope...
 
Mass_spec1.4906744_paper
Mass_spec1.4906744_paperMass_spec1.4906744_paper
Mass_spec1.4906744_paper
 
Identification of physical origin behind disorder, heterogeneity, and reconst...
Identification of physical origin behind disorder, heterogeneity, and reconst...Identification of physical origin behind disorder, heterogeneity, and reconst...
Identification of physical origin behind disorder, heterogeneity, and reconst...
 
Publication 2 (2014)
Publication 2 (2014)Publication 2 (2014)
Publication 2 (2014)
 
Multi-molecular views of a stellar nursery
Multi-molecular views of a stellar nurseryMulti-molecular views of a stellar nursery
Multi-molecular views of a stellar nursery
 
Symposium Final 2015
Symposium Final 2015Symposium Final 2015
Symposium Final 2015
 
ACS Applied & Mat Interfaces 2015
ACS Applied & Mat Interfaces 2015ACS Applied & Mat Interfaces 2015
ACS Applied & Mat Interfaces 2015
 
Non-parabolicity and band gap renormalisation in Si doped ZnO
Non-parabolicity and band gap renormalisation in Si doped ZnONon-parabolicity and band gap renormalisation in Si doped ZnO
Non-parabolicity and band gap renormalisation in Si doped ZnO
 
Assemblies of aziridinemethanols
Assemblies of aziridinemethanolsAssemblies of aziridinemethanols
Assemblies of aziridinemethanols
 
B.Zhao paper1-chem.mater-2005 SWNT bone mineralization
B.Zhao paper1-chem.mater-2005 SWNT bone mineralizationB.Zhao paper1-chem.mater-2005 SWNT bone mineralization
B.Zhao paper1-chem.mater-2005 SWNT bone mineralization
 
Energy-Dispersive X-Ray Spectroscopy Procedure for Analysing Cellular Element...
Energy-Dispersive X-Ray Spectroscopy Procedure for Analysing Cellular Element...Energy-Dispersive X-Ray Spectroscopy Procedure for Analysing Cellular Element...
Energy-Dispersive X-Ray Spectroscopy Procedure for Analysing Cellular Element...
 
Panushka UROP (1)
Panushka UROP (1)Panushka UROP (1)
Panushka UROP (1)
 

5. all published paper

  • 1. © 2010 WILEY-VCH Verlag GmbH & Co. KGaA, WeinheimAdv. Mater. 2010, 22, 3729–3734 3729 www.advmat.de www.MaterialsViews.com COMMUNICATION By Halei Zhai, Wenge Jiang, Jinhui Tao, Siyi Lin, Xiaobin Chu, Xurong Xu, and Ruikang Tang* Self-Assembled Organic–Inorganic Hybrid Elastic Crystal via Biomimetic Mineralization [*] H. Zhai, W. Jiang, J. Tao, S. Lin, X. Chu, Dr. X. Xu, Prof. R. Tang Center for Biomaterials and Biopathways and Department of Chemistry Zhejiang University Hangzhou, 310027 (P.R. China) E-mail: rtang@zju.edu.cn Dr. X. Xu, Prof. R. Tang State Key Laboratory of Silicon Materials Zhejiang University Hangzhou, 310027 (P.R. China) DOI: 10.1002/adma.201000941 It is generally accepted that biomaterials have unique physi- cochemical properties.[1] Inspired by biological systems, sci- entists have been studying biomimetic methods to fabricate functional materials.[2] Almost all biomaterials possess a common multi-component feature.[1,3] These composites fre- quently have ordered organic–inorganic hybrid structures and their properties are distinct from the individual components. For example, in a multilayered complex of inorganic aragonite tablets and organic substrate, the fracture toughness of nacre is significantly improved to three thousand times greater than that of synthetic aragonite.[4] Another striking example is biological bone. In bone, the hydroxyapatite (HAP) phase crystallizes in the nanoscaled channels formed by the staggered alignment of the protein matrix. The typical HAP crystals in bone are plate-shaped with extremely thin thickness (1.5–2 nm), which is the smallest known dimension of the biologically formed crystals.[5] In nature the organic and inorganic components inti- mately associate into well-organized hybrid structures to ensure optimal strength and flexural stress.[6,7] Therefore, in biomi- metic designs and fabrications the formation of such ordered nanostructures is a key challenge. The formation of inorganic crystals in living organisms is regulated by the organic matrix. Generally, different organic templates and additives lead to variety in the morphology, size, orientation, and assembly of the inorganic crystal by medi- ating its nucleation and growth.[8,9] Although many organic– inorganic nanocomposites have been reported,[10] the self- formation of ultrathin organic–inorganic substructures is still difficult to achieve by using a simple bottom-up approach. But the self-formed ordered and intimate combination of organic additives and inorganic crystals at the nanoscale is a crucial requirement for bioactive composites.[11] Here we prepare an organic–inorganic hybrid crystal by the self-assembly of cal- cium phosphate, surfactant, and protein. This hybrid crystal is composed of uniform and alternate organic–inorganic layers at the nanoscale. Both the inorganic crystalline phase and organic phases in the hybrid crystals have an ultrathin thick- ness of 1–2 nm. The two ordered components form simultane- ously during the crystal generation so that they integrate well with each other to form a superstructure. It is of great impor- tance that such biomimetic crystals are considerably flexible and elastic. It is believed that functional organic molecules can interact with calcium species at the organic–inorganic interfaces to modulate the growth and assembly process of the inorganic crystals. The globular protein bovine serum albumin (BSA), which comprises a single chain of 583 amino acid residues, is one of the most studied proteins. It is widely used as a model protein in many fields including biomimetic miner- alization.[12] Surfactants are widely applied as the crystalliza- tion templates in many biomimetic studies.[13] However, the cooperation of different organic additives has been frequently overlooked in previous works because of the complicated inter- actions in the system.[14] Actually, the interactions of a sur- factant molecule and protein are widely found in biological systems, for example, the interaction of protein with cell mem- brane surfactants. The selected two compounds can represent the protein matrix and special small functional molecules in biomimetic mineralization studies. Usually, proteins and sur- factants can form complexes in solution, which are frequently described by a “necklace bead model”. The micelle-like clus- ters of surfactants scatter along the polypeptide chains like the pearls in a necklace.[15] The hydrophilic groups of micelles are exposed to aqueous solutions and their configuration can be adjusted. In such protein–surfactant complexes, the protein is functionalized by the surfactant; meanwhile the aggrega- tion behavior of the surfactant is also affected by the protein structure. Here we find that the complex of BSA and an ani- onic surfactant (sodium bis(2-ethylhexyl) sulfosuccinate, AOT) could self-assemble into regular rhombus plates with a spe- cific organic–inorganic substructure in a calcium phosphate solution. Scanning electron microscopy (SEM) shows the uniform rhombic plates formed by the collaboration of calcium phos- phate, BSA, and AOT (Figure 1a).The typical rhombs are 300–400 nm in the long axis and 200–300 nm in the short axis. Their typical thickness is 80–100 nm. These rhombs are stable and their structures can endure in solution or in air for months. The energy-dispersive X-ray spectroscopy (EDS) reveals the pres- ence of calcium and phosphate ions in the rhombs; the atomic ratio of Ca:P is around 1.5. In addition to the elements of C and O, S was also detected (Figure S1 and Table S1 of the Sup- porting Information), indicating the presence of AOT (–SO3 2−). The organic–inorganic hybrid composite was also confirmed by
  • 2. 3730 www.advmat.de www.MaterialsViews.com © 2010 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim Adv. Mater. 2010, 22, 3729–3734 COMMUNICATION characteristic substructure: two independent sets of diffrac- tion peaks were detected by using wide-angle X-ray diffraction (WAXD) and small-angle X-ray diffraction (SAXD) (Figure 1e). In the small-angle region, a typical reflection characteristic of lamellar structures is observed. The interspacing distance, d = 3.12 nm, was calculated by using the reflection peak at 2θ = 2.83° ((001) reflection of the rhomb crystals). The (002) and (003) reflections were detected at 2θ = 5.71° (d = 1.55 nm) in SAXD and 2θ = 8.45° (d = 1.05 nm) in WAXD, respectively. These sharp peaks show the rhombs had a highly ordered lamellar structure. The other WAXD peaks in the normal range (2θ > 10°) indicate that the crystallized mineral phase is a HAP- like phase. These examinations clearly demonstrate that there are two independent lattice structures within a rhomb crystal. It is important that the organic and inorganic phases are orderly arranged to form the hybrid materials rather than the simple and disordered mixture. By using a side view of the ultrathin Fourier transform infrared spectroscopy (FTIR). The peaks at 1737, 1459, and 1419 cm−1 are the characteristic peaks of AOT, while the bands at 1655 (amide I) and 1553 cm−1 (amide II) indi- cate the presence of BSA. In addition, the broad peaks at 1023 and 567cm−1 areduetothepresenceoftheinorganicphosphategroup (Figure 1b). Thermogravimetric analysis (TGA) and differential scanning calorimetry (DSC) showed the presence of 21.4 % organic component (the organics decomposed at temperatures of 200–500 °C) and 62.1 % inorganic composite (the residue at temperatures above 500 °C, Figure 1c). From these results, we can conclude that the rhombs are the hybrid materials of inor- ganic (calcium phosphate) and organic phases (BSA and AOT). The regular rhombs were examined by means of trans- mission electron microscopy (TEM, Figure 1d). The selected area electron diffraction (SAED) pattern shows the inor- ganic phase is in a crystalline form and the pattern is similar to that of HAP tiny crystallites. Abnormally, the crystal has a Figure 1. a) SEM image of the rhombs. Inset: enlargement of the rhomb in the white circle. b) FTIR curves of the rhombs (bottom) and AOT (top). The characteristic peaks for BSA, AOT, and phosphate, are marked as circles, triangles, squares, respectively. c) TGA and DSC analysis under a nitrogen atmosphere. The weight percentages of water and organic component are labeled. d) Transmission electron microscopy (TEM) image of the rhombs. Inset: selected area electron diffraction (SAED) pattern corresponding to the white circled area. e) Wide-angle and small-angle (inset) X-ray diffraction (WAXD and SAXD, respectively) patterns of the rhombs. f) TEM side view of an ultrathin sectioned rhomb.
  • 3. 3731 www.advmat.de www.MaterialsViews.com © 2010 WILEY-VCH Verlag GmbH & Co. KGaA, WeinheimAdv. Mater. 2010, 22, 3729–3734 COMMUNICATION a strong binding effect with calcium ions as a result of the highly charged –SO3 2− groups (Figure S2, Supporting Information). But the interaction between calcium and BSA was rel- atively poor. Since AOT molecules aggregated onto the BSA chains according to the “neck- lace bead model”, the local concentrations of calcium around the BSA–AOT complex were greater than that in the bulk solution so that the AOT aggregates on BSA provided the het- erogeneous nucleation sites for calcium phos- phate. Moreover, the AOT molecules were organized by the BSA structure so that the complexes could induce the ordered assembly of calcium phosphate. We suggest that the mineral surfaces also act as the stable solid substrates for the self-assembly of the BSA– AOT complex. Thus, the lamellar organic– inorganic structures could be bottom-up assembled in the solutions spontaneously. Accordingly, the substructure of the hybrid rhombs is the alternate combination of the ultrathin nanocrystal layer and the BSA– AOT monolayer (Figure 2b), which is analo- gous to the nanoscale characteristics of many natural hybrid composites.[1,3,6,7] Structured materials are usually asso- ciated with unique physicochemical and biological properties.[16] Both advantages of inorganic and organic phases can be present in one hybrid material if these two components can be well-integrated at the nanoscale.[17] Although the main compo- nent of the rhombs is the crystallized cal- cium phosphate, a rigid inorganic phase, flexile and elastic behavior of the hybrid crystal was obtained. Figure 3a illustrates a side view of a rhomb: the whole crystal and its organic–inorganic layers are bent to some extent. Interestingly, a similar bent wave shape can be seen in the typical organic–inorganic hybrid reinforced materials such as some polymer–clay nanocoposites.[18] In order to confirm the mechanical features of the material, a force curve examination using atomic force microscopy (AFM, Figure 3b) was applied. The cantilever was very sensitive to the tip force and its deflection curve could qualitatively repre- sent the hardness of the examined surface. In contrast to the typical sudden and straight force–deflection lines for the rigid silicon substrate (modulus of 130 GPa, which is similar to that of pure HAP crystals: 112 GPa[19] ), the loading force increased smoothly with an increase of the deflection degree of the AFM cantilever. The buffer effect in the AFM force examina- tion indicates that the rhombs are not rigid. This characteristic was similar to that of a typical soft material, polystyrene (PS, modulus of about 3 GPa). It is interesting that no obvious per- manent damage or indention point was detected on the rhomb surface after the loading–unloading cycles (inset of Figure 3b) in the AFM examination. In order to quantitatively understand the mechanical properties of the hybrid, a nanoindentation measurement with a diamond indenter tip was additonally section of the rhombs under TEM, the lamellar structure is shown in Figure 1f: the dark region corresponds to the inor- ganic phase (crystallized calcium phosphate) and the light one is the organic phase. The individual organic and inorganic phases are alternately stacked. Each layer structure could be identified readily at the nanoscale in the hybrid crystal. These two distinct units are well integrated so that the complete hybrid crystals can be finally produced at the nanoscale. The thickness of each organic–inorganic unit is about 3.2 ± 0.2 nm, which is in good agreement with the calculated d value from the SAXD study. It is noted that the thickenss of the mineral layer is only about 2 nm; this dimension is close to that of biological ultrathin HAP crystallites formed between the collagen fibers of bone. In order to understand the substructure of the rhombs, the organic component was partially degraded by a 5 % NaOCl solu- tion. Thus, the mineral layer in the complex could be observed directly by TEM (Figure 2a). Small crystalline platelets, tens of nanometers in dimension (length and width), were frequently observed. In a rhomb crystal, the locations of inorganic crystal- line platelets are restricted by the adjacent protein–BSA organic frames. Thus, the continuous inorganic ultrathin layers might be formed between the frames by using the nanocrystallites. The conductivity investigations showed that AOT molecules had Figure 2. a) TEM image of the rhombs etched by 5 % NaOCl; The inset is its fast Fourier trans- form (FFT) image. b) Substructures of the organic–inorganic rhombs. AOT: small molecules with round head; BSA: long dark chains; mineral phase: rectangles.
  • 4. 3732 www.advmat.de www.MaterialsViews.com © 2010 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim Adv. Mater. 2010, 22, 3729–3734 COMMUNICATION could even partially recover during the unloading processes. In contrast, the unloading curves should be vertical if the solid phase was rigid.[20] Since the indentation depth was greater than 20 % of the sample thickness, the Bec model[21] for a thin soft material on a hard substrate was applied in the estimation of the modulus (see details in Supporting Information). By using the loading–unloading curves, the calculated modulus of the organic–inorganic rhombs was 6.64 ± 1.41 GPa. This value was even lower than the modulus of elastic-featured human vertebral trabeculae, 13.5 ± 2.0 GPa.[22] Similar to biological bone, both the elastic and hardness features were successfully integrated by the nanostructured assembly of organic and inor- ganic ultrathin phases, implying that the hybrid rhombs resolve the brittleness shortage of inorganic crystals and improve the material’s toughness. Actually, this is a smart and important strategy of living organisms to generate functional biomaterials by means of hybrid nanostructures. Many research efforts often focus on the controlling effect of the organic matrix on inorganic mineralization processes, which mediates the size, morphology, and orientation of inor- ganic crystals. Such an understanding implies a one-way con- trol of inorganic phase formation by organic additives. Thus, the organic templates are often required prior to the controlled crystallization in order to obtain hybrid materials. However, this understanding is not suitable in the current case. It was noted that the BSA–AOT complexes could not form the rhomb structure spontaneously in calcium solutions. Neither our experiment nor the published literature detected the BSA–AOT rhomb in the absence of any mineral phase. Only poorly crystal- line calcium phosphate spherical particles were obtained if only BSA was added into the calcium phosphate solution. Besides, AOT alone resulted in the conventional rod-like HAP crystals (Figure S3, Supporting Information) without any substructure. Clearly, the formation of the hybrid rhombs is attributed to the coexistence of BSA, AOT, and calcium phosphate, which is an emergent process. As mentioned above, the presence of the inorganic part also induces the assembly and structure of the organic components during mineralization.[23] Additionally, the changes of BSA and AOT concentrations within a certain range only affects the size and morphology of the resultant rhombs (Figure S4, Supporting Information). However, their internal substructure was not altered at all (Figure S5). This phenomenon could be explained by the regulation effect of surfactant on the complex assembly, which has been demon- strated by previous work.[15] We noted that the assembly process rather than conven- tional crystal growth occurred in the rhomb formation. No obvious signal between 50 and 100 nm was observed during the whole reaction process by dynamic light scattering (DLS, Figure 4). At the initial stage of crystallization, two individual distribution peaks existed in the DLS pattern. The small one (∼20 nm) represented the BSA–AOT building block in the reac- tion solution (Figure S6, Supporting Information), while the large one (∼300 nm) belonged to the final product. The frac- tion of the building block decreased gradually with the reaction, while the intensity of the product increased. Eventually, only the final product could be found at the end of the experiment. The product size did not increase during the reaction. Accord- ingly, the ex situ electron microscopy studies also demonstrated performed on the rhombs so that the modulus of the material could be calculated.[20] The solid and dashed lines represent the loading and unloading processes, respectively (Figure 3c). The relatively great indentation depth with different loading forces from 25 to 40 μN were used to demonstrate the elastic charac- teristic of the whole nanoplate well. Under such great external forces, the deformation of the plates was significant. However, the thin crystals did not collapsed and the depressed surfaces Figure 3. a) TEM image of ultrathin section of the rhomb. b) Atomic force microscopy (AFM) force curves of silicon substrate, rhombs (Rh) and poly- styrene (PS). Cantilever deflection represents the deformation distance of the sensitive AFM cantilever. Inset: AFM image of the plate after the loading–unloading cycle. c) The nanoindentation curves of rhombs. The displacement here means the indentation distance from the surface.
  • 5. 3733 www.advmat.de www.MaterialsViews.com © 2010 WILEY-VCH Verlag GmbH & Co. KGaA, WeinheimAdv. Mater. 2010, 22, 3729–3734 COMMUNICATION and nanoindentation were prepared by spin-coating 100 μL of slurry on silicon wafers (3000 rpm). For ultrathin-sectioned TEM examination, rhombs were embedded in 0.5 mL of epoxy. The mixture was solidified at 80 °C for 12 h and then carefully microtomed by a Reichert–Jung Ultracut E using a diamond knife. The typical thickness of the ultrathin sections was ∼80 nm. Characterization: SEM was performed by using a HITACHI S-4800 microscope at an accelerating voltage of 5 kV. FTIR spectroscopy (Nicolet Nexus 670) was used to determine the composition of the products. Thermogravimetric analysis was carried out by a TA Instrument SDT Q600. The experiment was measured in a temperature range of 22–600 °C at a heating rate of 10 °C min−1 under nitrogen atmosphere. TEM observations were performed by a Philips CM200UT microscope at a typical accelerating voltage of 160 kV. WAXD and SAXD were carried out by means of a Rigaku D/max-2550pc instrument with monochromatized Cu Kα radiation and a scanning step of 0.02°. AFM images were collected by a Veeco multimode scanning probe microscope with Nano IVa controller. The measurements were performed using an E head and a silica tip (Veeco) on a cantilever with a spring constant of 40 N m−1 in tapping mode with filters off, with a scanning rate of 20−60 Hz. The qualitative measurement of the mechanical properties was performed by the cantilever deflection in the AFM force curve. The data was collected for 200 individual force curves on 10 different rhombs. The nanoindentation measurements were performed by a Tribo-Indenter In-Situ Nanomechanical Test System with a Berkovich diamond indenter (tip radius of about 50 nm). The system was calibrated by using fused quartz before indentation. The data was collected using TestWorks 4 (MTS Systems). The modulus was calculated using the Oliver and Pharr method and the substrate effect was corrected by the Bec model. The DLS measurements were taken by using a Brookhaven Instruments 90 Plus particle size analyzer. Conductivity measurements were carried out by Conducometer DDS-11A at 30 °C. The conductivity electrode was calibrated using 0.01 M KCl solution prior to use. Supporting Information Supporting Information is available online from Wiley InterScience or from the author. Acknowledgements We thank Haihua Pan and Yuan Su for their helpful discussions, Yuewen Wang, Jieru Wang, Yin Xu, and Xiaoming Tang for assistance in material characterization techniques. This work was supported by the National Natural Science Foundation of China (20601023 and 20871102), Zhejiang Provincial Natural Science Foundation (R407087), the Fundamental Research Funds for the Central Universities and Daming Biomineralization Foundation. Received: March 16, 2010 Revised: April 5, 2010 Published online: July 21, 2010 the absence of intermediate solid or phase during the growth. The DLS result reveals an abnormal pathway in the organic– inorganic hybrid material assembly. We suppose that the BSA– AOT complexes induce the mineral crystallization firstly and then they are restructured by the mineral phase to form the alternative layer structure by a cooperative effect. However, the detailed mechanism needs further investigation. In this Communication we demonstrate that organic–inor- ganic hybrid rhombs with a lamellar superstructure can be self-generated by protein, surfactant molecules, and mineral phases. Each crystal contains two basic nanoscaled subunits: the ultrathin inorganic mineral and organic ultrathin layers. These layers are formed simultaneously and integrate well by self- assembly to generate the hybrid crystals. During this process the cooperative effect between the organic and inorganic phases is key. The ordered organic–inorganic nanostructure confers the optimum mechanical properties on the resultant hybrid mate- rial. The current study provides further evidence of the biomi- metic fabrication of functional materials. Experimental Section Materials: Triply distilled CO2-free water was used in the experiment. Ca(NO3)2 and (NH4)2HPO4 were of analytical grade and their solutions were filtered twice through Millipore films (0.22 μm) prior to use. BSA (Albumin Bovine fraction V, BR, purity >98 %) and AOT (Aldrich) were used directly without further purification. Preparation: An aqueous solution (100 mL) containing AOT (4 mM) and BSA (1 mg mL−1 ) was prepared. The solution pH was adjusted to 10.0 ± 0.5 at room temperature by ammonia solution (3 M). Ca(NO3)2 solution (50 mL, 5mM, pH = 10.0 ± 0.5) was added to the mixed solution at a rate of 10 mL min−1 and the solution was stired for 30 min. After that, (NH4)2HPO4 solution (50 mL, 5mM, pH = 10.0 ± 0.5) was added dropwise at a rate of 1.5 mL min−1. The slurry was examined by DLS periodically and the formed solids were collected by high-speed centrifugation at 10 000 rpm. All the solid samples were washed by water three times and were vacuum-dried at 35 ± 1 °C. Freshly prepared rhombs were dispersed in ethanol (∼0.5 mg mL−1 ) and collected on carbon-coated copper grids for TEM examination. Samples for AFM measurements Figure 4. Dynamic light scattering (DLS) size distribution curves at dif- ferent stages during the emergent formation of rhombs. The percentage values are calculated by using the statistics of the particle amounts. [1] S. Mann, Biomineralization: Principles and Concepts in Bioinorganic Materials Chemistry, Oxford University Press, Oxford 2001. [2] a) L. P. Lee, R. Szema, Science 2005, 310, 1148; b) C. Sanchez, H. Arribart, M. M. G. Guille, Nat. Mater. 2005, 4, 277; c) T. Kato, A. Sugawara, N. Hosoda, Adv. Mater. 2002, 14, 869; d) T. Sun, L. Feng, X. Gao, L. Jiang, Acc. Chem. Res. 2005, 38, 644. [3] a) N. Watabe, J. Ultrastruct. Res. 1965, 12, 351; b) L. C. Palmer, C. J. Newcomb, S. R. Kaltz, E. D. Spoerke, S. I. Stupp, Chem. Rev. 2008, 108, 4754; c) H. O. Fabritius, C. Sachs, P. R. Triguero, D. Roobe, Adv. Mater. 2009, 21, 391.
  • 6. 3734 www.advmat.de www.MaterialsViews.com © 2010 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim Adv. Mater. 2010, 22, 3729–3734 COMMUNICATION 2005, 285, 562; d) S. Chodankar, V. K. Aswal, P. A. Hassan, A. G. Wagh, Phys. B 2007, 398, 112; e) T. Chakraborty, I. Chakraborty, S. P. Moulik, S. Ghosh, Langmuir 2009, 25, 3062. [16] a) J. Aizenberg, A. Tkachenko, S. Weiner, L. Addadi, G. Hendler, Nature 2001, 412, 819; b) B. Pokroy, V. Demensky, E. Zolotoyabko, Adv. Funct. Mater. 2009, 19, 1054. [17] a) Z. Tang, N. A. Kotov, S. Magonov, B. Ozturk, Nat. Mater. 2003, 2, 413; b) L. J. Bonderer, A. R. Studart, L. J. Gauckler, Science 2008, 319, 1069. [18] a) P. Podsiadlo, A. K. Kaushik, E. M. Arruda, A. M. Waas, B. S. Shim, J. Xu, H. Nandivada, B. G. Pumplin, J. Lahann, A. Ramamoorthy, N. A. Kotov, Science 2007, 318, 80; b) M. A. Priolo, D. Gamboa, J. C. Grunlan, ACS Appl. Mater. Interfaces 2009, 2, 312. [19] G. Dewith, H. J. A. Vandijk, N. Hattu, K. Prijs, J. Mater. Sci. 1981, 16, 1592. [20] W. C. Oliver, G. M. Pharr, J. Mater. Res. 1992, 7, 1564. [21] a) S. Bec, A. Tonck, J. M. Georges, E. Georges, J. L. Loubet, Philos. Mag. A-Phys. Condens. Matter Struct. Defect Mech. Prop. 1996, 74, 1061; b) S. Roche, S. Bec, J. L. Loubet, in Mechanical Properties Derived from Nanostructuring Materials, Vol. 778 (Eds: D. F. Bahr, H. Kung, N. R. Moody, K. J. Wahl), Materials Research Society, Warrendale PA 2003, p. 117; c) G. Hochstetter, A. Jimenez, J. P. Cano, E. Felder, Tribol. Int. 2003, 36, 973. [22] J. Y. Rho, T. Y. Tsui, G. M. Pharr, Biomaterials 1997, 18, 1325. [23] M. Antonietti, M. Breulmann, C. G. Göltners, H. Cölfen, K. K. W. Wong, D. Walsh, S. Mann, Chem. Eur. J. 1998, 4, 2493. [4] a) J. D. Currey, Proc. R. Soc. London, Ser. B 1977, 196, 443; b) A. P. Jackson, J. F. V. Vincent, R. M. Turner, J. Mater. Sci. 1990, 25, 3173. [5] C. Burger, H. W. Zhou, H. Wang, I. Sics, B. S. Hsiao, B. Chu, L. Graham, M. J. Glimcher, Biophys. J. 2008, 95, 1985. [6] S. Weiner, H. D. Wagner, Annu. Rev. Mater. Sci. 1998, 28, 271. [7] D. Liu, H. D. Wagner, S. Weiner, J. Mater. Sci. Mater. Med. 2000, 11, 49. [8] F. C. Meldrum, H. Cölfen, Chem. Rev. 2008, 108, 4332. [9] a) G. Falini, S. Albeck, S. Weiner, L. Addadi, Science 1996, 271, 67; b) R. Kniep, S. Busch, Angew. Chem. Int. Ed. 1996, 35, 2624; c) J. Aizenberg, A. J. Black, G. M. Whitesides, Nature 1999, 398, 495; d) S. Sadasivan, D. Khushalani, S. Mann, Chem. Mater. 2005, 17, 2765. [10] a) S. Mann, Nat. Mater. 2009, 8, 781; b) R.-Q. Song, H. Cölfen, Adv. Mater. 2010, 22, 1301. [11] H. Gao, B. Ji, I. L. Jäger, E. Arzt, P. Fratzl, Proc. Natl. Acad. Sci. U. S. A. 2003, 100, 5597. [12] J. Xie, J. Y. Lee, D. I. C. Wang, J. Phy. Chem. C 2007, 111, 10226. [13] C. E. Fowler, M. Li, S. Mann, H. C. Margolis, J. Mater. Chem. 2005, 15, 3317. [14] a) L. Qi, J. Li, J. Ma, Adv. Mater. 2002, 14, 300; b) A. Kotachi, T. Miura, H. Imai, Chem. Mater. 2004, 16, 3191. [15] a) N. J. Turro, X. Lei, K. P. Ananthapadmanabhan, M. Aronson, Langmuir 1995, 11, 2525; b) C. K. Ober, G. Wegner, Adv. Mater. 1997, 9, 17; c) S. De, A. Girigoswami, S. Das, J. Colloid Interface Sci.
  • 7. Controlled formation of calcium-phosphate-based hybrid mesocrystals by organic–inorganic co-assembly Halei Zhai,a Xiaobin Chu,a Li Li,a Xurong Xuab and Ruikang Tang*ab Received 28th July 2010, Accepted 27th August 2010 DOI: 10.1039/c0nr00542h An understanding of controlled formation of biomimetic mesocrystals is of great importance in materials chemistry and engineering. Here we report that organic–inorganic hybrid plates and even mesocrystals can be conveniently synthesized using a one-pot reaction in a mixed system of protein (bovine serum albumin (BSA)), surfactant (sodium bis(2-ethylhexyl) sulfosuccinate (AOT)) and supersaturated calcium phosphate solution. The morphologies of calcium-phosphate-based products are analogous to the general inorganic crystals but they have abnormal and interesting substructures. The hybrids are constructed by the alternate stacking of organic layer (thickness of 1.31 nm) and well-crystallized inorganic mineral layer (thickness of 2.13 nm) at the nanoscale. Their morphologies (spindle, rhomboid and round) and sizes (200 nm–2 mm) can be tuned gradually by changing BSA, AOT and calcium phosphate concentrations. This modulation effect can be explained by a competition between the anisotropic and isotropic assembly of the ultrathin plate-like units. The anisotropic assembly confers mesocrystal characteristics on the hybrids while the round ones are the results of isotropic assembly. However, the basic lamellar organic–inorganic substructure remains unchanged during the hybrid formation, which is a key factor to ensure the self-assembly from molecule to micrometre scale. A morphological ternary diagram of BSA–AOT–calcium phosphate is used to describe this controlled formation process, providing a feasible strategy to prepare the required materials. This study highlights the cooperative effect of macromolecule (frame structure), small biomolecule (binding sites) and mineral phase (main component) on the generation and regulation of biomimetic hybrid mesocrystals. Introduction Scientists are eager to mimic nature’s ability to design functional materials whose properties are often superior to the synthetic ones. In nature, biominerals are widely produced by bacteria, protists, plants, invertebrates and vertebrates, including humankind.1 These biological materials are featured by a smart combination of multi-components especially in the form of integrated organic–inorganic hybrid materials, in which the organic parts are often proteins and low-molecular-mass mole- cules.2 They are constructed by using organic components to control the nucleation, growth, organization and transformation of inorganic phases. Interactions between organic and inorganic phases at the molecular level, although complex, are common occurrences to determine the size, shape, and properties of the resulting products.1,3 Different from the synthesized ones, the functions of biominerals depend to a large extent on the ordered association of biomolecules with mineral phases. The organized hybrid materials, unlike the single components, can be tailored into different compositions and morphologies, e.g. bone,4 tooth5 and mollusc shells6 etc., to ensure the optimal mechanical and physicochemical characteristics. The controls that determine the sizes, shapes, and properties of crystals are a key to addressing numerous challenges in material designs and applications. It has been revealed that organic molecules can influence the shape and properties of inorganic crystals.7 However, it is difficult for the two distinct organic and inorganic phases to spontaneously assemble into highly ordered structures. In living organisms, biological mineralization is able to combine particular building blocks or entities into functional hybrid composites. An understanding of these biochemical controls is essential and important, not only to study biomineralization mechanisms further, but also to design novel hybrid materials and processing technologies. Despite the complicated hierarchical structures of biominerals, their basic building blocks are frequently the nano-sized organic–inorganic composites.8 Therefore, an ordered and periodic assembly of organic and inorganic nanophases at the nanoscale is crucial to biomimetically synthesize hybrid materials. But, how can we design ordered hybrid composites and how can we conveniently control their structures, sizes and morphologies under mild conditions? Although organic–inorganic hybrid materials have been approached by various methods such as layer-by-layer (LbL)9 and template-directed crystallization,10 the bottom-up fabrica- tion from ions or molecules is still a great challenge in the laboratory since the control of periodic deposition is difficult to achieve at hierarchical scales. In conventional biomimetic crys- tallization studies, organic molecules, which act as structure- directing agents, modulate the crystal morphology by their a Centre for Biomaterials and Biopathways, Zhejiang University, Hangzhou, Zhejiang, 310027, China. E-mail: rtang@zju.edu.cn; Fax: +86-571-87953736; Tel: +86-571-87953736 b State Key Laboratory of Silicon Materials, Zhejiang University, Hangzhou, Zhejiang, 310027, China 2456 | Nanoscale, 2010, 2, 2456–2462 This journal is ª The Royal Society of Chemistry 2010 PAPER www.rsc.org/nanoscale | Nanoscale Publishedon13October2010.Downloadedon24/01/201605:58:33. View Article Online / Journal Homepage / Table of Contents for this issue
  • 8. selective absorption onto crystal faces, altering crystal facet stability and growth kinetics.7,11 Recently, a non-classical crystal growth pathway based upon nano assembly has received considerable attention.12 The nanoparticles, which are directed by specific organic additives, can act as the basic building units to assemble into superstructures or mesocrystals. During such a process, the organic molecules (especially macromolecules) selectively absorb and interact with primary nanocrystals. The assembly process follows programmed arrangement into high order hybrid structures.13 The morphology can be tuned by varying the interactions between different organic and inorganic phases. However, the one step bottom-up process, which starts from the molecular level rather than from preformed nano- particle precursors, may be readily able to control the orientation and order of assembly processes to form integrated hybrid nanocomposite. But this strategy requires a precise and sponta- neous co-assembly of both organic and inorganic phases alternately at both the molecular level and the nanoscale.14 In this paper, we reported an easy but effective method for direct synthesis of organic–inorganic hybrid mesocrystals by a emergent co-assembly process of protein (bovine serum albumin (BSA)) and surfactant (sodium bis(2-ethylhexyl) sulfo- succinate (AOT)) in a supersaturated calcium phosphate solu- tion. The calcium-phosphate-based hybrid crystals with lamellar structure have different properties from conventional ones. Here we emphasize that the size and morphology of the resulting hybrids could be regulated readily by varying BSA, AOT and calcium phosphate concentrations according to a suggested morphological ternary diagram. This study provided a novel pathway to one-pot preparation of functional hybrid crystal materials with tuneable size and morphologies by organic–inor- ganic co-assembly. Results and discussion It is believed that functional organic molecules can interact with calcium species at the organic–inorganic interfaces to modulate the growth and assemble of inorganic crystals. BSA is one of the most studied proteins but this biological macromolecule is not an effective modifier in calcium phosphate crystallization.15 It has been previously confirmed that the interaction between BSA and calcium or phosphate ions in aqueous solutions is poor.16 BSA itself is inert in mineral deposition. In contrast, many surfactant molecules are widely used as effective promoters and templates in biomimetic calcium mineralization since their hydrophilic groups (especially the sulfonate and carboxylate groups) provide active binding sites to calcium ions. AOT is one among typical agents that can modulate calcium phosphate precipitation significantly. AOT molecules have a strong binding effect with calcium ions due to their highly charged -SO3 2À groups.16,17 However, hierar- chical or complicated biomineral-like structures cannot be achieved by using this small molecule due to the lack of higher- order structures. In our control experiments, only poor crystal- line HAP was obtained if BSA was added into the supersaturated calcium phosphate solutions; AOT alone produced the conven- tional rod-like HAP crystals without any organized hybrid structure. These results matched the previous studies and understandings well. However, the cooperative effect of BSA and AOT in the calcium phosphate solution could lead to the formation of unique hybrids in a one-pot reaction. Under an experimental condition of 2 mM AOT, 1 mg mlÀ1 BSA and 1.25 mM calcium ions (the molar ratio of calcium to phosphate was fixed at 1.67 in all experiments), the uniform rhombic plates precipitated spontaneously as shown by scanning electron microscopy (SEM, Fig. 1(A)). Their size distribution was homogeneous. The typical rhombic plates were 1.23 Æ 0.21 and 0.91 Æ 0.18 mm along their long and short axes, respectively (statistical results from $100 plates); the aspect ratio was about 1.4. The thickness of the plates was 130 Æ 20 nm. These rhombic plates had exactly same morphology (Fig. 1(B)) and this char- acteristic was similar to the general inorganic crystals. However, the chemical compositions of the obtained plates were relatively complicated. Besides the elements of calcium and phosphorus, the element of sulfur was detected in the solids by using energy- dispersive X-ray spectroscopy (EDS). This result indicated the presence of AOT (-SO3 2À ) in the hybrid plates. It was also revealed that inorganic part in the plates was a kind of calcium phosphate minerals with Ca : P molar ratio of 1.5–1.6. The coexistence of organic–inorganic components was also confirmed by Fourier transform infrared spectroscopy (FT-IR, Fig. 1(C)). The peaks at 1737, 1459 and 1419 cmÀ1 were the characteristic signals of AOT, while the bands at 1656 (amide I) and 1555 cmÀ1 (amide II) showed the involvement of BSA in the solids.18 The broad peaks at 1022 and 564 cmÀ1 were assigned to the inorganic phosphate groups.19 Thermogravimetric analysis (TGA) showed that the mineral phase was the main composition in the solids. The weight loss of 38% between 100 and 500 C was corresponded predominantly to removal of the organic phase, while the weight contents of the inorganic phases were 62%. In addition, the plates became ‘crimped-paper’-like after calcina- tions at 500 C in air for 2 h. Without the organic frame, the solids became brittle and the structures were collapsed readily into small pieces under an ultrasonic condition. Many previous studies suggested that the organic compounds play a regulation role in inorganic mineralization rather than being involved in Fig. 1 (A) SEM image of the rhombic plates. (B) Enlarged image of the rhombic plate in the white circle; the double-headed arrow shows the extended orientation. (C) FT-IR pattern of the products. (D) The rhombic plates after calcination at 500 C in air. This journal is ª The Royal Society of Chemistry 2010 Nanoscale, 2010, 2, 2456–2462 | 2457 Publishedon13October2010.Downloadedon24/01/201605:58:33. View Article Online
  • 9. structural recombination. However, the current results implied that BSA and AOT were the key components in the hybrid construction. Thus, these solids were different from the other precipitated inorganic crystals in the presence of organic addi- tives. The resulting rhombic plates shared the same size and aniso- tropic morphology similar as general inorganic crystals. However, in-depth examination revealed that they were distinct from the conventional calcium phosphate crystals.20 The rhombic plates were examined by wide angle X-ray diffraction (WAXD, Fig. 2(A)) and as expectated, the crystalline HAP-like calcium phosphate phase was detected. The WAXD pattern was very similar to that of pure HAP but small peak shifts were also observed. We suggested that the binding effect between the organic component and calcium ions would cause the lattice distortion. The lattice structure of the inorganic phase could be revealed at the atomic scale by using high resolution transmission electron microscopy from a top view of the plates (HRTEM, Fig. 2(B)). This image represents a typical ultrathin inorganic crystal layer embedded in the rhombic plates. However, another independent set of diffraction peaks was found in the X-ray diffraction (XRD) pattern, which revealed that a superstructure was present in the hybrids. The characteristic peaks of lamellar structure (interspacing distance, d ¼ 3.43 nm) could be found from both small angle X-ray diffraction (SAXD) and WAXD (Fig. 2(A)), indicating an ordered arrangement of subunits along a crystallographic direction rather than a simple mixture of the organic and inorganic phases. A side view of the ultra-thin sectioned samples under transmission electron microscopy (TEM, Fig. 2(C)) confirmed the internal structure: the organic layers (light, 1.31 nm) and the inorganic layers (dark, 2.13 nm) alternately stacked at the nanoscale to form the compact hybrid structure. Thus, the organic molecules (BSA and AOT) were well organized to form the layered organic phase. Each organic– inorganic ultra-thin unit had a thickness of 3.44 nm, which agreed with the XRD data, 3.43 nm, and the individual inorganic layer was a calcium phosphate crystal plate with a thickness of only 2.13 nm. These nanoplates acted as the building blocks that could self-assemble together with the organic layers to generate the lamellar complex. Additionally, a wave-like superficial texture of the hybrids could be observed (Fig. 1(B)) and the profiles were similar to the hybrid crystal morphology. This phenomenon indicated that the assembly might be an anisotropic process. In order to understand the orientation of each inorganic layer, selected area electron diffraction study (SAED, Fig. 2(D)) was applied. It was noted that the anisotropic diffraction dots rather than the isotropic diffraction rings were obtained during the examination of a whole rhombic crystal, which represented a similar characteristic of single crystal. It was interesting that the orientation reflected by these dots (arrow in the insert image) was exactly same as the long axis of the examined rhombic crystal. Such a coincidence implied that all the ultrathin inorganic crystal layers within the hybrid plates should share the same crystallo- graphic orientation. Additionally, the experimental diffractions dots of the whole crystal were almost same as the fast Fourier transform (FFT) result (Fig. 2(B)) of an individual crystal layer. Therefore, the formed hybrid crystal exhibited similar features to a single crystal; however, it had additional superlattice structure. Since the rhombic plates had a specific morphology while they were not constructed as the conventional single crystals, these hybrids could be considered as a kind of artificial meso- crystal.12,21 However, the imperfect dots on Fig. 2 (D) might indicate that the misaligned orientation still occurred during nano assembly. Since the material was constructed by ultrathin calcium phosphate units, it was interesting that flexible and elastic features were conferred onto the mesocrystal along the lamellar packing direction in spite of that; its main composition was a brittle ceramic phase. These mechanical properties of the hybrids had been characterized by our previous study,16 demonstrating the advantages of organized assembly for formation of mesocrystals in material functionalization. The convenient control of the size and morphology of the organic–inorganic hybrids and mesocrystals is a challenge, although those for single hybrid crystals are nowadays sophis- ticated. In our experiments, the calcium phosphate–BSA–AOT hybrid mesocrystals with different size and morphology could be feasibly regulated within a simple reaction system by changing the reactant concentrations. We fixed BSA and calcium concentration at 0.5 mg mlÀ1 and 1.25 mM, respectively. When the AOT concentration was 1.00 mM, the obtained hybrid plates were not rhombic plates any more. Their shapes became spindle- like. The hybrid plates changed into a round shape when the AOT concentration was increased to 4.00 mM. However, the further decreasing or increasing of AOT concentration result into the disappearance of the co-assembly or hybrid in the system. In this experiment, their morphologies were gradually adjustable from spindle, to rhombus to round by increasing the AOT concentration from 1.00 to 4.00 mM (Fig. 3). During the evolution process, the length along the short axis of the formed Fig. 2 (A) WAXD and SAXD (insert) patterns of the rhombi; (B) HRTEM of a rhombus (top view). Insert: FFT simulation result; (C) TEM image of ultra-thin sectioned rhomb from side view. The values of 2.13, 1.31 and 3.44 nm corresponded to the thicknesses of inorganic (dark), organic (light) and organic–inorganic complex layers, respec- tively. Insert: TEM image of the side view of the ultra-thin sections of the plates, bar is 0.5mm. (C) is the enlargement of the region within the white circle; (D) TEM image of the hybrids. Insert was the SEAD pattern (white circle area). The HRTEM image in (B) was also obtained on the same area by the in situ technique. Arrows showed that each individual inorganic plate in the hybrid shared the same crystallographic orienta- tion, which was the long axis of the rhombus. 2458 | Nanoscale, 2010, 2, 2456–2462 This journal is ª The Royal Society of Chemistry 2010 Publishedon13October2010.Downloadedon24/01/201605:58:33. View Article Online
  • 10. hybrid plates did not change significantly, it was maintained at 300–400 nm. However, the long axis kept on decreasing from 1.50 mm to 300–400 nm with increasing AOT concentration. Accordingly, the hybrid morphology became isotropic. This phenomenon implied that AOT component was an important factor to control the a degree of anisotropic co-assembly of the hybrids. Although the morphologies and sizes of the resulted hybrid plates were influenced remarkably by the changing of AOT concentrations in the reaction solutions, the internal organic– inorganic subunit remained. The WAXD and SAXD patterns of the spindles, rhombi and rounds were exactly same without any change. But the misalignments of each individual inorganic layer in the hybrid increased with the increasing of AOT concentra- tion. The crystallographic mismatch of the inorganic layers could be examined by using SEAD. During the evolution from the regular rhombi to round shapes, the diffraction dots disappeared gradually while the diffraction rings existed (Fig. 4). This tendency indicated that the preferred orientation of the thin calcium phosphate planes in the hybrid was weakened. Although AOT itself could result in aggregates in solution to induce calcium mineralization, the aggregation was simple and isotropic due to the lack of complicated configuration. Therefore, it was reasonable that the excessive AOT could destroy the anisotropic assembly of the ultrathin mineral plates in the hybrid rhombi. Although the inorganic and organic layers were still packed layer-by-layer strictly along the thickness direction, the crystal- lographic directions of the inorganic crystal planes in the hybrids became disordered. The anisotropic assembly transformed the orientation of the long axes into the isotropic mode with increasing AOT concentration; thus, the round plates were finally yielded at 4.00 mM AOT and the hybrid was not meso- crystalline any more. Besides, it should be mentioned that the percentages of organic and inorganic contents in the hybrid solids was not changed significantly during the morphology modulation; in which the inorganic content was kept within a range of 69–72% from the spindles to the rounds. Besides the AOT concentrations, the formation of hybrid crystals could be also adjustable by BSA concentration. In this examination, the concentrations of AOT and calcium were maintained at 2 mM and 1.25 mM, respectively, and the BSA concentrations were increased from 0.25 to 2.00 mg mlÀ1 . It was noted that the morphologies of hybrid plates underwent another gradual evolution from the irregular quadrilaterals to rhombi and then to plump spindles (Fig. 5). The sizes and aspect ratios of the hybrids increased from 200 nm to 2 mm and 1.1 to 2.0, respectively, during the modulation. Although the hybrid width increased along the short axis, the more extended length along the long axis indicated that the anisotropy assembly process was affected significantly by the protein concentration. It was noticed that in biomineralization, the complicated hierarchical building structures of biominerals are frequently contributed by the ordered aggregates of proteins. Again, the basic organic–inor- ganic units and their ordered packing behaviours were not changed during the morphology and size regulations. It was mentioned that, when the BSA concentration increased, the role of AOT in the synthesis decreased. Therefore, the ratio or the Fig. 3 SEM images of the hybrids synthesised at AOT concentrations of 1.00 (A), 2.00 (B) and 4.00 mM (C). (D)–(F) are the corresponding XRD patterns of (A)–(C), respectively. Fig. 4 During the morphology change from rhombus (A) to round (B), anisotropic diffraction dots became isotropic rings in the corresponding SEAD pattern. Fig. 5 SEM images of the hybrids at BSA concentration of 0.25 (A), 1.13 (B) and 2.00 mg mlÀ1 (C). This journal is ª The Royal Society of Chemistry 2010 Nanoscale, 2010, 2, 2456–2462 | 2459 Publishedon13October2010.Downloadedon24/01/201605:58:33. View Article Online
  • 11. cooperative effect of BSA and AOT was another key factor in mesocrystal formation and regulation. It was known that the co-assembly could not occur in the absence of the inorganic phase. Thereby, it was reasonable that the concentrations of calcium and phosphate could control the mesocrystal formation too (Fig. 6). Under BSA and AOT concentrations of 1.00 mg mlÀ1 and 2.00 mM, respectively, the resulting rhombi shared the same intermediate state with increasing calcium and phosphate concentration in the reaction solution. If calcium concentration was decreased to 0.63 mM, the poly-dispersed quadrilaterals-like plates (size of 400–800 nm) formed with the small aspect ratio of 1.1. If the concentration was increased to 2.50 mM, the slender spindle-like plates were obtained and their size distribution was 1.8–2.3 mm with an aspect ratio was 2.5. From the evolution from quadrilaterals, rhombi to slender spindles, it could be seen that the anisotropic co-assembly process was enhanced. The previous studies of biomimetic fabrication of hybrid materials with artificial molecules such as peptide-amphiphile,22 block copolymer,23 and amphiphilic dendro-calixarene,24 sug- gested that the specific sites and sterically constrained effect may control the assembly of the organic template and then the size and morphology of the final hybrid materials. Different from the above-mentioned understanding, under our experimental conditions, the change of BSA and AOT concentrations were directly related to the different modification state of BSA. The BSA protein, which was constituted by a single chain of 583 amino acid residues, acted as a stable and relatively rigid fragment connected with the special motif (AOT aggregates).25 The hydrophilic groups of aggregates exposed to aqueous solu- tions and their configuration can be adjusted. The highly charged group (–SO3 2À ) in AOT could greatly interact with calcium ions and then modulated calcium phosphate precipitation significantly, which had been demonstrated experimentally in many works and in our previous paper.16,26 However, the binding ability of BSA with calcium ions is weak and the controlling effect on the mineral formation is relatively poor. As a result, BSA acted as structural frame while the AOT aggregates provided the nucleation sites of mineral during the co-assembly process. In the current study, BSA macromolcules combined with smaller AOT molecules to form a BSA–AOT complex and such a modified protein could effective control the crysallization and assembly of the calcium phosphate mineral. To some extent, this method provides an efficient way to turn a non-mineraliza- tion protein into a mineralization protein by using surfactants. The conformation of the macromolecules restricts the assembly only along certain specific directions. However, the larger concentration of AOT is accompanied by an increase in the amount and size of AOT aggregates, offering more sites for the assembly process.27 As a result, the controlling effect from the protein was counteracted and the assembly process could happen at more directions to form the isotropic rounds. Furthermore, increasing the amount or the relative amount of BSA concen- trations partly restricted the assembly process in specific prefer- ential orientations by spatial configuration to form the anisotropic hybrids or mesocrystals.28 Thus, the co-assembly process preferred to occur in certain directions, especially along the long axis of the hybrid plates rather than the short axis. Although the short axis partly extended under some experi- mental cases, the greatly increase along the long axis resulted into the spindles-like mesocrystal formation. The competitive controlling effect of BSA and AOT led to the transformation of an isotropic and anisotropic assembly process during hybrid crystal construction. Thus, the formation of different hybrids and mesocrystals with tuneable size and morphologies could be achieved. An anisotropic co-assembly process could also be promoted by increasing the mineral ion concentrations. In the formation process of mesocrystals, the inorganic precursor controlled the size and morphology of the final product by tuning the amounts, size and shapes of the nano-sized building blocks.29 Under our experimental conditions, the controlling role of mesocrystal growth became dominant in greater saturation to decide the product size and structure. As the preferred orientation of the calcium phosphate crystal plates is parallel to the long axis of the rhombic plates, the fast growth of the calcium phosphate plate crystals along this preferred orientation promoted the formation of the slender spindle-like plates with larger aspect ratios during the co-assembly process. However, the interaction between BSA-AOT complex and calcium phosphate crystal was also responsible for the co-assembly of the organic and inorganic phase to form highly ordered hybrid materials and maintain their internal structure. Actually, the generation of hybrid material via the cooperative effect of macromolecules (mainly proteins), small biomolecules and the mineral phase is a common strategy in natural bio- mineralization.30 In the biological construction, high-molecular- weight macromolecules, such as collagen, act as support matrix to provide a structural frame for the mineralization, the biomineralization proteins themselves have nucleation sites but most matrices receive mineralization function by binding and stabilizing functional motifs that are carboxylate- or Fig. 6 SEM images of the hybrids at calcium concentration of 0.63 (A), 1.56 (B) and 2.50 mM. In all experiments, the ratio of calcium to phos- phate in the reaction solution was maintained at 1.67. 2460 | Nanoscale, 2010, 2, 2456–2462 This journal is ª The Royal Society of Chemistry 2010 Publishedon13October2010.Downloadedon24/01/201605:58:33. View Article Online
  • 12. sulfonate-rich. Thus, the combination of organic–inorganic mineralization interfaces and the organized organic matrices can concentrate the mineral ions to induce the deposition as well as to regulate the size, morphology and orientation of the inorganic building blocks to form integrated organic–inorganic hybrid composites with complicated structure. We suggest that in this system, BSA is the structural frame to control the anisotropic assembly; the adsorption of AOT onto BSA enhances the mineralization ability of the protein; and the mineral acts as an inorganic conjunction phase to solidify the organic–inorganic hybrid structure. In the experiments, the increase of BSA promoted the formation of larger hybrid plates with increased aspect ratio, while AOT exhibited the opposite controlling effect. The increasing of inorganic concentrations preferred the formation of slender hybrid plates with a larger size. In order to show the controlling effect of the reactant concentrations, the simplified morphological maps in the form of solution ternary diagrams was proposed (Fig. 7). The biomimetic formation hybrid and mesocrystals could be yielded in the grey region. In the specific regions, the formed hybrid plates had a similar size and morphology. From points A to B, the increase of aspect ratio was preferred as the hybrid rounds transformed into the spindle ones. Since the anisotropic assembly behaviour was enhanced, this evolution implied that the resulting mesocrystals became more organized and the mismatch degrees of the inorganic layers in the hybrids could be reduced. From points A to C, both the size and aspect ratio of the resulted hybrids were increased and their morphologies were changed from rounds to spindles. From points B to C, the hybrids turned from wide spindles to slender spindles with increased size and aspect ratio too. By using this morphological ternary diagram, we could design readily hybrids and mesocrystals with the required size and morphology. Conclusions We demonstrate that the ordered and uniform hybrids or mes- ocrystals can be biomimetically synthesized by the co-assembly of proteins, small functional molecules and minerals using a simple one-pot reaction. Their size distributions and morphologies can be adjusted by varying the component concentration in reaction solutions. The anisotropic co-assembly of the BSA–AOT complex and ultrathin calcium phosphate crystal plates is a key to the control of mesocrystal formation. A morphological ternary diagram can be used to design different hybrid materials as requireed. This work may give another inspiration to the assembly of multi components into one inte- grated hybrid material with a highly ordered structure. Furthermore, the bottom-up pathway of controlled fabrication may be developed as a simple and effective strategy to prepare feasibly functional hybrid and mesocrystal materials. Experimental Materials Triply distilled water was used in all the experiments. Ca(NO3)2 and (NH4)2HPO4 were of analytical and their solution were filtered twice using 0.22mm Millipore films prior to use. BSA (Albumin Bovine fraction V, BR, purity 98%, LABMAX) and AOT (Aldrich) were used without any further purification. Hybrid plate preparation Using a typical experiment as an example, 100 ml aqueous solution containing 4 mM AOT and 0.20 g BSA was mixed with 50 ml Ca(NO3)2 solution (5mM). The solution pH was adjusted to 10.0 Æ 0.5 at room temperature by 3 M ammonia solution. Then 50 ml (NH4)2HPO4 solution (3 mM, pH ¼ 10.0 Æ 0.5) was added dropwise at a rate of 1.5 ml minÀ1 . The reaction solution contained 2.00 mM AOT, 1.00 mg mlÀ1 BSA, 1.25 mM Ca(NO3)2 and 0.75 mM (NH4)2HPO4. The mixture was gently stirred at 30 Æ 1 C for 24 h. The precipitated solids were collected by centrifugation at 6000 rpm. The solid were washed by water for three times and were vacuum-dried at 35 Æ 1 C. In order to examine the controlling effect of reactant concentrations on hybrid formation, different concentrations of AOT, BSA and calcium phosphate ions were used and all the experimental processes were the same. Fig. 7 Controlled synthesis of hybrids by a morphological ternary diagram. The co-assembly occurred within the grey area and the formation of mesocrystals was preferred in its left and bottom sections. The typical morphology of the final products were also demonstrated. Bar ¼ 1mm. This journal is ª The Royal Society of Chemistry 2010 Nanoscale, 2010, 2, 2456–2462 | 2461 Publishedon13October2010.Downloadedon24/01/201605:58:33. View Article Online
  • 13. Characterizations SEM was performed by using a HITACHI S-4800 at a typical acceleration voltage of 5 kV. FT-IR spectra (Nicolet Nexus 670) were applied to analysis the hybrid compositions. WAXD and SAXD were characterized by a Rigaku D/max-2550pc with monochromatized Cu-Ka radiation; the scanning step was 0.02 . TGA was performed by a TA Instrument SDT Q600. The experiment was measured in a temperature range from room temperature to 1000 C under nitrogen atmosphere. TEM observations were performed by a CM200UT TEM (Philips) at an acceleration voltage of 160 kV. During the ultra-thin sectioned TEM examination, rhombi were embedded in epoxy. The mixture was solidified at 80 C for 12 h and then carefully microtomed by a Reichert-Jung Ultracut E using a diamond knife. Acknowledgements We thank Jieru Wang, Xinting Cong, Xiaomin Tang, Yin Xu and Linshen Chen for their help with characterization, Haihua Pan and Yuan Su for discussions. This work was supported by the Fundamental Research Funds for the Central Universities, National Natural Science Foundation of China (20871102), Zhejiang Provincial Natural Science Foundation (R407087) and Daming Biomineralization Foundation. Notes and references 1 S. Mann, Biomineralization: Principles and Concepts in Bioinorganic Materials Chemistry, Oxford University Press, 2001. 2 L. Bedouet, F. Rusconi, M. Rousseau, D. Duplat, A. Marie, L. Dubost, K. Le Ny, S. Berland, J. Peduzzi and E. Lopez, Comp. Biochem. Physiol., Part B: Biochem. Mol. Biol., 2006, 144, 532–543; J. L. Arias and M. a. S. Fernacndez, Chem. Rev., 2008, 108, 4475–4482. 3 C. E. Killian and F. H. Wilt, Chem. Rev., 2008, 108, 4463–4474; J. S. Evans, Chem. Rev., 2008, 108, 4455–4462. 4 S. Weiner and H. D. Wagner, Annu. Rev. Mater. Sci., 1998, 28, 271–298. 5 S. Busch, U. Schwarz and R. Kniep, Chem. Mater., 2001, 13, 3260–3271. 6 N. Watabe, J. Ultrastruct. Res., 1965, 12, 351–370. 7 F. C. Meldrum and H. C€olfen, Chem. Rev., 2008, 108, 4332–4432. 8 R. Z. Wang, Z. Suo, A. G. Evans, N. Yao and I. A. Aksay, J. Mater. Res., 2001, 16, 2485–2493; H. J. Gao, B. H. Ji, I. L. Jager, E. Arzt and P. Fratzl, Proc. Natl. Acad. Sci. U. S. A., 2003, 100, 5597–5600. 9 Z. Tang, N. A. Kotov, S. Magonov and B. Ozturk, Nat. Mater., 2003, 2, 413–418; P. Podsiadlo, A. K. Kaushik, E. M. Arruda, A. M. Waas, B. S. Shim, J. Xu, H. Nandivada, B. G. Pumplin, J. Lahann, A. Ramamoorthy and N. A. Kotov, Science, 2007, 318, 80–83. 10 N. Gehrke, N. Nassif, N. Pinna, M. Antonietti, H. S. Gupta and H. C€olfen, Chem. Mater., 2005, 17, 6514–6516; P. H. Kithva, L. Grondahl, R. Kumar, D. Martin and M. Trau, Nanoscale, 2009, 1, 229–232. 11 N. A. J. M. Sommerdijk and G. d. With, Chem. Rev., 2008, 108, 4499–4550. 12 R. Q. Song and H. C€olfen, Adv. Mater., 2010, 22, 1301–1330. 13 M. Li, H. C€olfen and S. Mann, J. Mater. Chem., 2004, 14, 2269–2276. 14 S. Mann, Nat. Mater., 2009, 8, 781–792. 15 R. I. Martin and P. W. Brown, J. Mater. Sci.: Mater. Med., 1994, 5, 96–102; K. L. Yadav and P. W. Brown, J. Biomed. Mater. Res., 2003, 65a, 158–163. 16 H. Zhai, W. Jiang, J. Tao, S. Lin, X. Chu, X. Xu and R. Tang, Adv. Mater., 2010, 22, 3729–3734. 17 C. E. Fowler, M. Li, S. Mann and H. C. Margolis, J. Mater. Chem., 2005, 15, 3317–3325. 18 G. Falini, S. Weiner and L. Addadi, Calcif. Tissue Int., 2003, 72, 548–554. 19 S. J. Gadaleta, E. P. Paschalis, F. Betts, R. Mendelsohn and A. L. Boskey, Calcif. Tissue Int., 1996, 58, 9–16. 20 J. Song, V. Malathong and C. R. Bertozzi, J. Am. Chem. Soc., 2005, 127, 3366–3372; A. Ethirajan, U. Ziener, A. Chuvilin, U. Kaiser, H. C€olfen and K. Landfester, Adv. Funct. Mater., 2008, 18, 2221–2227; Y. Zhang and J. Lu, Cryst. Growth Des., 2008, 8, 2101–2107. 21 A.-W. Xu, M. Antonietti, S.-H. Yu and H. C€olfen, Adv. Mater., 2008, 20, 1333–1338. 22 J. D. Hartgerink, E. Beniash and S. I. Stupp, Science, 2001, 294, 1684–1688; V. M. Yuwono and J. D. Hartgerink, Langmuir, 2007, 23, 5033–5038. 23 Z. H. Nie, D. Fava, E. Kumacheva, S. Zou, G. C. Walker and M. Rubinstein, Nat. Mater., 2007, 6, 609–614; H. Wang, A. J. Patil, K. Liu, S. Petrov, S. Mann, M. A. Winnik and I. Manners, Adv. Mater., 2009, 21, 1805–1808. 24 M. Kellermann, W. Bauer, A. Hirsch, B. Schade, K. Ludwig and C. B€ottcher, Angew. Chem., Int. Ed., 2004, 43, 2959–2962. 25 N. J. Turro, X. G. Lei, K. P. Ananthapadmanabhan and M. Aronson, Langmuir, 1995, 11, 2525–2533; S. De, A. Girigoswami and S. Das, J. Colloid Interface Sci., 2005, 285, 562–573. 26 S. Sarda, M. Heughebaert and A. Lebugle, Chem. Mater., 1999, 11, 2722–2727. 27 C. K. Ober and G. Wegner, Adv. Mater., 1997, 9, 17–31. 28 H.-A. Klok, J. F. Langenwalter and S. Lecommandoux, Macromolecules, 2000, 33, 7819–7826; X. Kong and S. A. Jenekhe, Macromolecules, 2004, 37, 8180–8183; L. Rubatat, X. Kong, S. A. Jenekhe, J. Ruokolainen, M. Hojeij and R. Mezzenga, Macromolecules, 2008, 41, 1846–1852; A. Sacnchez-Ferrer and R. Mezzenga, Macromolecules, 2010, 43, 1093–1100. 29 H. C€olfen and M. Antonietti, Angew. Chem., Int. Ed., 2005, 44, 5576–5591. 30 N. Kroger, R. Deutzmann, C. Bergsdorf and M. Sumper, Proc. Natl. Acad. Sci. U. S. A., 2000, 97, 14133–14138; L. C. Palmer, C. J. Newcomb, S. R. Kaltz, E. D. Spoerke and S. I. Stupp, Chem. Rev., 2008, 108, 4754–4783. 2462 | Nanoscale, 2010, 2, 2456–2462 This journal is ª The Royal Society of Chemistry 2010 Publishedon13October2010.Downloadedon24/01/201605:58:33. View Article Online
  • 14. Spontaneously amplified homochiral organic–inorganic nano-helix complexes via self-proliferation† Halei Zhai,a Yan Quan,a Li Li,a Xiang-Yang Liu,b Xurong Xuc and Ruikang Tang*ac Most spiral coiled biomaterials in nature, such as gastropod shells, are homochiral, and the favoured chiral feature can be precisely inherited. This inspired us that selected material structures, including chirality, could be specifically replicated into the self-similar populations; however, a physicochemical understanding of the material-based heritage is unknown. We study the homochirality by using calcium phosphate mineralization in the presence of racemic amphiphilic molecules and biological protein. The organic–inorganic hybrid materials with spiral coiling characteristics are produced at the nanoscale. The resulted helixes are chiral with the left- and right-handed characteristics, which are agglomerated hierarchically to from clusters and networks. It is interesting that each cluster or network is homochiral so that the enantiomorphs can be separated readily. Actually, each homochiral architecture is evolved from an original chiral helix, demonstrating the heritage of the matrix chirality during the material proliferation under a racemic condition. By using the Ginzburg–Landaue expression we find that the chiral recognition in the organic–inorganic hybrid formation may be determined by a spontaneous chiral separation and immobilization of asymmetric amphiphilic molecules on the mineral surface, which transferred the structural information from the mother matrix to the descendants by an energetic control. This study shows how biomolecules guide the selective amplification of chiral materials via spontaneous self-replication. Such a strategy can be applied generally in the design and production of artificial materials with self-similar structure characteristics. 1 Introduction Through long time periods of evolution, most spiral coiled bio- materials in nature, like gastropod shells, adopt a specic homochirality.1,2 For example, the majority of current gastropod shells have a right-handed (R-) coiling pattern (Fig. 1a).3 The chiral minority was eliminated eventually by a frequency- dependent selection and the dominant one proliferated.4,5 This inspires us that materials with specic structural properties, like chirality, can spontaneously develop into a large self-similar community.6 Biologically, it is accepted that regularly expressed biomolecules, together with inorganic minerals, constitute the physical chirality of gastropod offsprings under the guidance of a controlling gene.7 For instance, at the growth front of shells, the tiny chitin nanocrystals behave as the amphiphilic mole- cules and self-assemble into the liquid crystal layers (Fig. 1b).8 Fig. 1 The chirality of gastropod shells and a schematic drawing of the shell mineralization front. (a) General gastropod species have right-handed shells. (b) During the natural generation of shell structure b-chitin molecules assemble into supermolecules (chitin crystallites) and their liquid-crystal layers induce the spiral mineralization of calcium carbonate (this scheme is prepared based upon a mechanism proposed by Cartwright et al.).8,9 a Centre for Biomaterials and Biopathways and Department of Chemistry, Zhejiang University, Hangzhou, 310027, China. E-mail: rtang@zju.edu.cn; Fax: +86 571- 8795-3736 b Department of Physics and Department of Chemistry, National University of Singapore, Singapore 117542, Singapore c Qiushi Academy for Advanced Studies, Zhejiang University, Hangzhou, Zhejiang 310027, China † Electronic supplementary information (ESI) available: Supporting gures and tables. See DOI: 10.1039/c3nr33782k Cite this: Nanoscale, 2013, 5, 3006 Received 23rd November 2012 Accepted 29th January 2013 DOI: 10.1039/c3nr33782k www.rsc.org/nanoscale 3006 | Nanoscale, 2013, 5, 3006–3012 This journal is ª The Royal Society of Chemistry 2013 Nanoscale PAPER Publishedon06February2013.DownloadedbyUniversityofCalifornia-SanFranciscoon24/01/201605:50:31. View Article Online View Journal | View Issue
  • 15. These chitin layers provide growing sites for the inorganic phase and modulate the mineralization together with related proteins. Thus, it follows that the spiral micro-pattern consti- tuted by a chitin–calcium carbonate lamellar structure is grad- ually constructed (Fig. 1b).9 In this sense, an understanding on the physicochemical regulations of the organic–inorganic bio- inspired materials with selective chirality will advance our knowledge in chemistry and materials sciences. A challenging question to be addressed is whether we can mimic the self- evolution (symmetry breaking) process of shells in our labora- tories so that the chiral materials can be separated and propa- gated to generate self-similar articial production. It has been demonstrated that some organic molecules can control the morphology of biominerals, like calcium phosphate and calcium carbonate crystals.10 The chiral organic molecules usually act as templates to control the crystal morphology, rather than incorporated organic composition to constitute the chiral hybrid materials.11,12 In the articial design of chiral nanomaterials, a variety of dispersed chiral superstructures, such as nano-helixes and nano-tubes, can be generated with the twisted assembly of chiral molecules, or even nano-sized crys- talline units.13 However, each nano-helix or nano-tube is con- structed by independent assembly, rather than a successive proliferation procedure to pass down the chirality and nal formation of the homochiral complex. As a result, the archi- tecture of a homochiral material complex is rarely achieved.14 Herein, by employing a racemic mixture of a chiral amphiphile (bis-(2-ethylhexyl) sulfosuccinate sodium salt, AOT) and bovine serum albumin (BSA) in supersaturated calcium phosphate solution, two kinds of chiral organic–inorganic hybrid nano- helixes (L- and R-enantiomers) can spontaneously form and each kind of chiral helix eventually proliferates into a larger homochiral helix complex. We feel that such an experimental phenomenon may be relevant to the proliferation of chiral materials. 2 Experimental section 2.1 Materials Triply distilled CO2-free water was used in the experiment. Ca(NO3)2 and (NH4)2HPO4 were of analytical grade and their solutions were ltered twice through 0.22 mm Millipore lms prior to use. BSA (Albumin Bovine fraction V, BR, purity 98%) and AOT (Aldrich, racemic mixture) were directly used without further purication. 2.2 Preparation of the homochiral nano-helix complex The temperature during all the synthesis processes was main- tained at 30 Æ 1 C. Briey, a 100 ml aqueous solution con- taining 1 mM AOT and 1 mg mlÀ1 BSA was prepared. The solution pH was adjusted to 10.0 Æ 0.5 by 3 M ammonia solu- tion. 50 ml Ca(NO3)2 solution (5 mM, pH ¼ 10.0 Æ 0.5) was added. Then 50 ml (NH4)2HPO4 solution (3 mM, pH ¼ 10.0 Æ 0.5) was added dropwise at a rate of 1.5 ml minÀ1 . The solution was gently stirred for 10 h and the formed solids were collected by centrifugation at 3600 rpm. All the solid samples were washed by water three times and were vacuum-dried at 35 Æ 1 C. Freshly prepared samples were dispersed in ethanol ($0.5 mg mlÀ1 ) and collected on carbon-coated copper grids for TEM examinations. In the seed growth experiment, 1/20 percent of the obtained product underwent intense ultrasonic treat- ment (KUDOS, 35 kHz, 20 min) and the helix clusters or networks were collapsed into dispersed helixes. Then the dispersed helixes were added as seeds into freshly prepared reaction solutions and the reaction solutions were collected by centrifugation and observed with Transmission Electron Microscopy (TEM). For ultrathin sectioned TEM examination, dried samples were embedded in 0.5 ml epoxy. The mixture was solidied at 80 C for 12 h and then carefully microtomed by a Reichert-Jung Ultracut E ultramicrotome using a diamond knife. 2.3 Au-labelled BSA absorptions BSA-Au nanoparticles were synthesized according to the work by J. Xie et al.15 Briey, 5 ml 10 mM HAuCl4 solution was added into 5 ml 50 mg mlÀ1 BSA solutions and stirred for 5 min. Then, 0.5 ml 1 M NaOH was added and the solution was kept at 37 C for 24 h. The product was dialyzed with 1000 ml distilled water for 24 h. The BSA-Au was used instead of pure BSA in order to probe the location of BSA on the surface of the helix. 2.4 Examination of calcium concentrations The concentration of free calcium ions in BSA, AOT, BSA + AOT solutions were measured by a PCa-1 calcium ion selective electrode with a saturated calomel electrode as the reference electrode. The electrode was calibrated according to the instructions before use. 2.5 Characterizations Scanning electron microscopy (SEM) was performed by using a HITACHI S-4800 eld-emission scanning electron microscope at an acceleration voltage of 5 kV. Fourier-transform infrared spectroscopy (FT-IR, Nicolet Nexus 670) was used to determine the composition of the products. Thermogravimetric analysis (TGA) was carried out by a TA Instrument SDT Q600. The experiments were measured over a temperature range of 22– 800 C at a rate of 10 C minÀ1 under air atmosphere. TEM observations were performed by a JEM-1200EX at a typical acceleration voltage of 80 kV. Small angle X-ray diffraction (SAXRD) and Wide angle X-ray diffraction (WAXRD) were char- acterized by a Rigaku D/max-2550pc with monochromatized Cu Ka radiation and the scanning step was 0.02 . Solid state nuclear magnetic resonance (ssNMR) was kindly performed by Prof. Jarry Chan's group at the National Taiwan University on a Bruker DSX300 NMR spectrometer. 3 Results and discussion 3.1 Structure and composition of the nano-helix In our biomimetic case, AOT and BSA were adopted as the models for biological amphiphilic and proteins, respectively. AOT is of asymmetric double-chain amphiphile (Fig. S1†). It can This journal is ª The Royal Society of Chemistry 2013 Nanoscale, 2013, 5, 3006–3012 | 3007 Paper Nanoscale Publishedon06February2013.DownloadedbyUniversityofCalifornia-SanFranciscoon24/01/201605:50:31. View Article Online
  • 16. assemble into various mesomorphous phases, which have been widely used in biomimetic crystallization.16 BSA is one of the common proteins in biomineralization studies.17 The syner- gistic effect of AOT and BSA on calcium phosphate minerali- zation gave rise to the formation of nano-helix (Fig. 2 and S2†). In the control experiments, the use of AOT and BSA alone only generated the calcium phosphate nanorods and nanospheres, respectively (Fig. S3†). Clearly, the helix formation was attrib- uted to the coexistence of AOT, BSA and calcium phosphate. It followed that the individual helixes could further develop into micron-sized aggregated clusters (Fig. 2a and b). In an individual cluster, the nano helixes extended radially outward from a dense core, indicating the successive proliferation procedure. Furthermore, some clusters connected with each other to form a larger network (Fig. 2 and S4†). As the basic building blocks of the clusters, the helixes were chiral and they had two kinds of spiral enantiomers, L- and R-forms. Although the overall amounts of the L- and R-helixes in the reaction system were equal (Fig. S5†), only one helix enantiomer could be identied within a cluster or connected network (Fig. 2a–c, see more in ESI†). This suggested that the spontaneous chiral recognition and chiral separation occurred during the cluster and network generation. Concerning the composition and structure of the helixes, they were constituted by organic and mineral phases, which accounted for 20.5 wt% and 58.8 wt%, respectively (the rest 20.7% was attributed to absorbed and crystal water, Fig. S6†). FT-IR (Fig. 2d) and energy-dispersive X-ray spectroscopy (EDS, Fig. S7†) revealed that the main components in the helixes were AOT and a calcium phosphate phase. X-ray diffraction (Fig. 2e) showed that the mineral phase was close to brushite. Moreover, the mineral phase in the helix was conrmed by Multiple Pulse Sequence Nuclear Magnetic Resonance spectroscopy (CRAMPS-NMR) and a Heteronuclear Correlation (HETCOR) spectrum between the 31 P and 1 H nuclei 31 P{1 H} combined rotation, indicating that the phosphate groups were protonated (HPO4 2À ) in the calcium mineral (Fig. 2f, S8 and Table S1†). The NMR data indicated the absence of PO4 3À in the complex. As a result, the calcium phosphate species containing PO4 3À groups such as hydroxyapatite, octa- calcium phosphate and tri-calcium phosphate could be excluded in the phase analysis. Additionally, Ca(H2PO4)2 could not be precipitated under our experimental conditions due to its high solubility. Both examinations shows that the signals of the helix were close to those of brushite. Therefore, the brush- ite-like mineral was considered as the primary inorganic component in the helixes. The internal structure of the nano-helixes could be consid- ered as the alternative and spiral stacking of thin calcium phosphate phase and AOT bilayers. The cross-section images of the nano-helix showed that the thickness of the wall of the nano-helix was about 2.1–2.2 nm (Fig. 3a–c). Furthermore, SAXRD and WAXRD results also showed the alternative lamellar superstructure in the helixes with a constant interspacing distance (d ¼ 3.34 nm). The lamellar structure was also demonstrated by TEM (Fig. 2e): the dark lines (1.7 nm) and light lines (1.6 nm) correspond to the inorganic calcium phosphate and organic AOT ultrathin layers, respectively. The thickness of each organic–inorganic hybrid unit was about 3.3 Æ 0.2 nm, which is in agreement with the d value calculated from the SAXRD data. In the spiral helix, there existed a pitch angle of about 43 between the strip edge and the long axis. It was noted that the AOT molecules preferred to assemble into a bilayer structure. Concisely, the organic bilayer could have a thickness of about 1.6 nm if the molecules tilted by 43 . There were two mirror forms for both the helix pitch angle and the AOT tilt angle, +43 or À43 , as the denitions (Fig. 3d and e). The mirror packing of AOT corresponded to the formation of R- and L-enantiomers of the helixes. Apart from AOT, a small amount of BSA was detected in the helix by FT-IR and 13 C{1 H} NMR (Fig. S9†). Using nano Au particle labelled BSA as the imaging agent, we found that the protein did not incorporate into the hybrid inner structure, but absorbed onto the helix wall surfaces (Fig. S10†), which might be due to its relatively large dimension.19 We suggested that BSA served as a surface or Fig. 2 Characterizations of nanohelixes. (a and b) Homochiral clusters consisting of R-helixes and L-helixes, respectively. (c) A homochiral helix network; circles indicate the cluster centres; inset is a magnification of the rectangular region. (d) FT-IR of the helix and pure AOT. The typical and undisturbed peaks of AOT (1750 cmÀ1 ), BSA (1540 cmÀ1 , amino) and phosphate ions could be noted. The peak located at 2342 cmÀ1 was generally attributed to CO2 from the air during the FT-IR determination. (e) SAXRD and WAXRD patterns of the helixes. d ¼ 3.34 nm and d ¼ 1.65 nm represent the first and the second diffractions of the lamellar structure in the helixes, respectively. WAXRD also showed that the mineral phase was similar to brushite. The XRD peaks of 11.3 and 31.0 were close to those of brushite (020) and (121), respectively. The case of the small left-shift of charac- teristic peaks could be found in small nanocrystals.18 The inset TEM image shows each organic (light line)–inorganic (dark line) unit in the helix. (f) 31 P{1 H}HETCOR spectra between the 31 P and 1 H nuclei measured in the helixes. The spectra was acquired at a spinning frequency of 10 kHz and the contact time was set to 2.5 ms. A total of 64 transients with an increment of 100 ms was accumulated. 3008 | Nanoscale, 2013, 5, 3006–3012 This journal is ª The Royal Society of Chemistry 2013 Nanoscale Paper Publishedon06February2013.DownloadedbyUniversityofCalifornia-SanFranciscoon24/01/201605:50:31. View Article Online
  • 17. structure stabilizer for the hybrid spiral strips. An experimental fact was that no chiral product formed by using BSA alone. During biomineralization, some biomacromolecules can adopt an extended conformation when they interact with the inor- ganic phase surface.20 Therefore, there was a possibility that the BSA molecules might incorporate into the helix using their extended forms. However, we could conrm that the AOT molecules with highly charged sulphuric groups, rather than BSA, were the primary organic composition in the nano-helix. For example, the FT-IR study showed a very weak amino peak (Fig. 2d) in the composites, implying the ignorable contents of BSA in comparison with the strong AOT bands. Therefore, it was suggested that the hybrid helixes were formed by the ordered assembly of the calcium phosphate mineralized layer and the AOT bilayer, and were stabilized by BSA absorption on the hybrid surfaces. In this architecture, the assembly behaviour of chiral AOT molecules in the hybrid helixes determined the material’s chirality. 3.2 Proliferation of the nano-helixes Originally, the homochiral clusters and networks evolved from a single nano-helix (mother matrix). The preformed chiral nano- helix spontaneously passed down the structure information (chirality) from one generation to the next and then generated the homochiral complexes (Fig. 4). Firstly, tiny hybrid buds sprouted from the surface of the matrix (Fig. 4a). Both organic and inorganic parts of the buds directly integrated with the corresponding parts of mother matrix. This could be considered as a kind of matrix outgrowth. Secondly, these hybrid buds grew longer and twisted into the helical ribbons. At this stage, the organic and inorganic parts at the growing front of hybrid buds did not integrate with mother matrix any more. At this time, there should be a choice in the twist direction (L- or R-). Nevertheless, we noted that the newly formed helical ribbons replicated precisely the twist direction (chirality) of the mother matrix. This meant that the mother matrix induced the later AOT molecules to assemble into a coherent packing direction, even the AOT bilayers at the budding region and growth front were separated by calcium phosphate layers. Thus, the original structure was inherited through the budding and proliferation process (Fig. 4b). Thirdly, the homochiral proliferation process of the helixes continued by generating more “daughters” and “grand-daughters” based upon the matrix. Due to the space limitation, the newly formed helixes tended to stretch outward, which generated radial homochiral clusters (Fig. 4c). Finally, a few of the helixes at the cluster edge acted as “bridges” to provide additional growing sites for new buds and initiated another proliferation process (Fig. 4e). This new proliferation Fig. 3 (a and b) Cross-section images of nano-helixes under TEM. (c) Schematic structure of the nano helixes (dark grey: inorganic phase; light grey: organic phase). (d and e) TEM and schemes of the R- and L-helix. The width of the AOT bilayer is 1.6 nm from TEM observation. As AOT molecules have a length of 1.1 nm, AOT molecules in a bilayer should arrange with a tilt angle of about 43 . Note: the AOT molecules in the same bilayer are simply treated as direct contact and this small variation of tilt angle doesn't affect our qualitative analysis. Fig. 4 TEM images of the evolution from a single helix to a homochiral cluster and then a homochiral network (community). (a) A sprouting bud from R-helix matrix for the new “daughter” helix generation. (b) Growth and twist of the “daughter” helix, which duplicated the chiral feature to be R-form; insets show the details of the growth front on the matrix. (c) More buds formed and they replicated the structure of matrix precisely. (d) Rudiment of the homochiral helix cluster; insets: magnification of the branching sites. (e) Homochiral helix cluster (R-form); arrows indicate the proliferation directions of the cluster; inset shows the new buds formed at an extended helix. (f) Homochiral helix networks (R- form); arrows show the proliferation directions. This journal is ª The Royal Society of Chemistry 2013 Nanoscale, 2013, 5, 3006–3012 | 3009 Paper Nanoscale Publishedon06February2013.DownloadedbyUniversityofCalifornia-SanFranciscoon24/01/201605:50:31. View Article Online
  • 18. could happen at multiple directions (Fig. 4f). Through the self- repeating processes, a single nano-sized helix eventually evolved into a large homochiral complex (network) at the micrometer scale (Fig. 4f). In each network, the chirality of the newly born helixes was precisely “inherited” from the original mother matrix from generation to generation, which can be considered as a spontaneous process of material-based self-proliferation. We note that it is impossible for the dispersed helixes aer intense ultrasonic treatment to aggregate into homochiral clusters again. However, aer the dispersed helixes are re- dispersed into the freshly prepared reaction solutions as seeds, the time to induce the formation of helix clusters can be rela- tively reduced according to the seed amount, indicating that the mother helix acts as the seed to induce the proliferation of new helixes to form helix clusters (Fig. S11†). As a result, in each network, the chirality of the newly born helixes was precisely “inherited” from the original mother matrix from generation to generation, which can be considered as a spontaneous process of material-based self-proliferation. 3.3 Model for the homochiral material Analogous to the shell formation, the co-assembly of organic and inorganic phases is restricted at the local domains of the growth front (Fig. 4a and b). AOT molecules are more able of binding calcium ions than BSA (Table S2†) and these amphiphilic molecules greatly modulate the growth of calcium phosphate species.21 In this case, the helix formation and replication are controlled dominantly by the assembly behaviours of AOT at the growth front. The AOT molecules can tightly absorb onto the surface of calcium phosphate species with a strong binding effect between calcium ions and sulphuric groups, which facili- tates the assembly of AOT bilayers.22 Unlike the free AOT mole- cules in aqueous solution, the relatively rigid CaP crystal, rather than the mobile water layer, can imobilize the adjacent AOT bilayers. Thus, the hybrid structure could be ‘solidied’ and stabilized with a decreasing disordered uctuation of AOT molecules comparing the free state, which then induces the next layer mineralization.23 This alternative and cooperated deposi- tion of the AOT bilayer and calcium phosphate phase layer gradually constitutes a thin AOT–calcium phosphate hybrid strip, which is similar to the associated assembly of lipids and inorganic phase reported by Seddon, et.al.13b In our work, the chiral AOT molecules are responsible for the twist of the hybrid strip to form the L- and R-nano-helixes. Although the BSA used here is constituted with chiral L-amino acids, the chirality of nano-helixes are unlikely to be controlled by BSA. Due to its large size, it is difficult to incorporate into the ordered structure with ultra-small units of 1.7 nm (calcium phosphate layer) and 1.6 nm (organic layer), while the twisted arrangement of these units forms the chirality at the nanoscale. In addition, the equal number of L- and R-nano-helixes also indicates that the BSA with single chiral units (L-amino acids) has little contribution to the chirality of the nano-helixes. Many works have been reported that strips constituted with chiral molecules tend to twist into nano-helixes to reduce the elastic energy.24 Similarly, the chirality of the nano-helixes in our system is determined by the assembly behaviour of chiral AOT molecules. However, the racemic mixture generally dilutes the chiral interaction between the chiral molecules, so that the chiral superstructure might fail to form.25 Nevertheless, some studies have shown that both R- and L-enantiomers can emerge in racemic systems if an energy favoured chiral phase separation occurs, especially for the lipids with chiral headgroups and inexible double chains structures.26 Interestingly, AOT owns a similar structure and phase behaviour to these lipids.27 More- over, chiral molecules can also undergo a phase separation when they are restricted at interfaces.28 Therefore, in our system, it follows that a spontaneous chiral phase separation of amphiphilic AOT may occur on the calcium phosphate mineral substrate, resulting in the bilayers with exactly the same molecular packing behaviour. Due to the complicated structure, the conformation infor- mation (chirality) of AOT in each nano-sized helix is difficult to identify. Besides, methods of the synthesis or separation of AOT diastereoisomers is rarely reported.29 Based upon the mirror arrangement of AOT in L- and R-helixes, we divide the AOT molecules into two types with different tilt directions of +43 or À43 . Aer this simplication, only the tilt angle needs to be taken into account in the qualitative analysis of the energy during the formation process. The AOT molecules in the bila- yers have two different tilt angles, +43 or À43 , which can be considered as the enantiomers to induce R- and L-chiral helix formations, respectively (Fig. 3d and e). The favoured tilt angle should maintain the same value during the alternative dispo- sition. Thus, the energy favoured recognition is a key to main- tain the molecular assembly according to the chiral breaking model supposed by Selinger et al.30 The model suggests that the elastic energy of the strip can be reduced by a chiral separation even under racemic conditions. An order parameter, j, is introduced, which is treated as the local net amount of right- handed minus le-handed molecular packing here. The elastic free energy, F, of the thin chiral bilayer strip can be written as eqn (1) F ¼ ð dS 1 2 k 1 r 2 þ 1 2 k0 1 r 2 cos2 f À lHPj 1 r sin f cos f þ 1 2 KðVjÞ2 þ 1 2 tj2 þ 1 4 uj4 þ Eedge (1) where, S is the area, the rst term is the standard Helfrich bending energy of the hybrid membrane and the coefficient k is the isotropic rigidity. In the right side of eqn (1), the second term represents the anisotropy of the rigidity and the coefficient k0 is the anisotropic term and f is the title angle of the chiral molecules (Fig. 5a); the third term is a chiral term that favours twisting in a tilt angle f; the coefficient lHP, is the chirality parameter, which exists only in chiral membranes and depends on the chiral order. The sign of lHP can be changed when the membrane transforms into its mirror image. lHPj increases with the greater chiral phase separation degree of j. The last three terms in the bracket are the Ginzburg–Landau expres- sions in powers of j, which represent the free energy change 3010 | Nanoscale, 2013, 5, 3006–3012 This journal is ª The Royal Society of Chemistry 2013 Nanoscale Paper Publishedon06February2013.DownloadedbyUniversityofCalifornia-SanFranciscoon24/01/201605:50:31. View Article Online
  • 19. during the ordering transition. The values of K and u are temperature independent constants while the coefficient of t relates to temperature and t 0 for chiral phase separation.30,31 Because AOT molecules in the helix have the xed tilt angle of 43 , the domain wall energy on the edge is a constant. When simultaneously minimizing the free energy over tilt angle and radius, r, in eqn (1), the following is obtained, f0 ¼ arctan k þ k0 k 1= 4 (2) r0 ¼ k 1= 4 ðk þ k0 Þ 1= 4 h k 1= 2 þ ðk þ k0 Þ 1= 2 i lHPj (3) in which, (k + k0 )/k represents the energy cost for the ratio of the bend parallel to the tilt direction to bend perpendicular to the tilt direction. In our system, its value is about 0.76 because fAOT ¼ 43 , which indicates that the hybrid strip favours twisting parallel to the AOT tilt direction. Besides, the radius of the nano-helix equals 1.33k/lHPj. Usually, the lipid amphi- philes form helical tubes or a helix with a larger diameter of hundreds nanometres or even a few micrometers. Since the organic–inorganic hybrid structure exists in our helixes, it is reasonable that the rigidity coefficient k should be greater than single component chiral lipid membranes. As a result, lHPj must be signicantly greater to produce such a slender helix with a very small radius ($10 to 25 nm). We note that the favoured energy barrier DF plays an essential role to control the twist direction and chiral prolifer- ation of nano-helixes (eqn (1)). Here, the qualitative description energy barrier DF between the racemic state (j ¼ 0) and sepa- ration state (jmÆ) can be described by eqn (4),30 DF ¼ ð dS lHPjmÆ 1 r sin f cos f þ 1 2 KðVjmÆÞ2 þ 1 2 tjmÆ 2 þ 1 4 ujmÆ 4 (4) As the radius of the nano-helix is r f 1/lHPj, the relatively small radius of the helix (10–25 nm) implies that lHPj is of great value in our case, which facilitates the chiral separation. The equation shows that the free energy of the strip has two local minima representing the two types of energy favoured AOT packing with the mirror symmetry (Fig. 5b) if chiral phase separation occurs. Without chiral phase separation, the strip cannot twist into a helix because the radius becomes innite when j ¼ 0 (r / N). Fig. 5b shows the constant arrangement of AOT molecules (constant tilt angle of +43 or À43 ) within a strip is energeti- cally preferred due to an energy barrier. First, the energy barrier DF can promote the formation of energy favoured and stable nano-helixes, rather than unstable hybrid strips. If the different arrangements of AOT molecules (L- and R-) coexist in the same strip, the elastic energy increases so that the resulting strip becomes unstable (Fig. 5b, middle). Therefore, the elastic energy cannot be reduced and generate twisted nano-helixes. By contrast, the same arrangements of AOT molecules (L- or R-) can successfully reduce the unfavoured elastic energy and twist to form L- or R-nano-helixes, respectively (Fig. 5b, le and right). Second, the energy barrier of DF is also responsible for the homochiral proliferation. In our system, the preformed helix matrix has an inductive effect on the sequent proliferation because the emerging organic and inorganic parts in the new buds directly extend from their mother matrix. Thus, new buds share the same AOT packing form with the mother matrix. The same AOT packing can be replicated under the guidance of the mother matrix due to the favoured energy reduction, which means that L- to L- or R- to R-proliferation is a preferential way. Subsequently, the buds grow following the determined AOT packing to form a new chiral helix with the same chirality. For example, the new buds generated from the R-nano-helixes in Fig. 4 faithfully adopt the R-twist direction and keep the selected form during the growth process. The mutated proliferation of L- to R- or R- to L- also require extra energy to overcome DF in comparison with the matched L- to L- or R- to R-. Accordingly, the chiral structure proliferation always initiates at the pre- formed helixes and amplies the chiral structure from the mother matrix to subsequent generations. Finally, large homochiral complexes (helix clusters and networks) can be generated under the guidance of the energy controlled recog- nition of AOT packing. 4 Conclusions This study reveals that the homochiral complex of the organic– inorganic hybrid helix can form via a self-proliferation process. The energy controlled chiral recognitions and separations of asymmetric chiral AOT molecules are essential in both helix formation and homochiral proliferation. The nding is of importance to approach homochiral biomimetic materials in the laboratory. We expect this strategy of bio-inspired chiral structure proliferation can be developed into a convenient pathway for the articial synthesis of self-similar functional materials. Acknowledgements We thank Prof. Jerry Chen for the ssNMR studies, Dr Jinhui Tao, Dr Haihua Pan and Yuan Su for discussions, Hua Wang, Jieru Fig. 5 (a) The geometry of AOT molecules in the helix discussed in eqn (1) for helix formation. (b) Two local minima of the elastic free energy (F) with symmetry packing (jm+ or jmÀ) lead to an energy barrier of DF, which ensures the oriented packing vector of AOT bilayers to produce chiral helix and homochiral proliferation. This journal is ª The Royal Society of Chemistry 2013 Nanoscale, 2013, 5, 3006–3012 | 3011 Paper Nanoscale Publishedon06February2013.DownloadedbyUniversityofCalifornia-SanFranciscoon24/01/201605:50:31. View Article Online