You are on page 1of 7

Langmuir 2006, 22, 575-581 575

Molecular Assembly of Metallacarboranes in Water: Light Scattering


and Microscopy Study
Pavel Matějı́ček,*,† Petr Cı́gler,‡,§ Karel Procházka,† and Vladimı́r Král‡
Department of Physical and Macromolecular Chemistry, Charles UniVersity in Prague,
HlaVoVa 2030, 128 40 Prague 2, Department of Analytical Chemistry, Institute of Chemical Technology,
Technicka 5, 166 28 Prague 6, and Institute of Organic Chemistry and Biochemistry, Academy of Science
of the Czech Republic, FlemingoVo n.2, 166 10 Praha 6, Czech Republic

ReceiVed August 12, 2005. In Final Form: October 27, 2005

Polyhedral metallacarboranes are used mainly as ion-pairing agents and recently have been recognized as potent
inhibitors of HIV protease. They are characterized by exceptional hydrophobicity, rigid geometry, delocalized negative
charge, ion-pairing behavior, and strong acidity of their conjugated acids. The completely novel phenomenon, association
of these promising pharmaceutical tectons in aqueous solutions, is described here. The behavior of two structural types
of metallacarboranes, [bis(1,2-dicarbollide)cobaltate(1-)] and bis[(3)-1,2-dicarbollylcobalt]-(3,6)-1,2-dicarbacanastide-
(2-)], in aqueous solution was studied by a combination of static and dynamic light scattering and microscopy
methods. Spherical aggregates with radii of ca. 100 nm and fairly monodisperse nanostructures were found in aqueous
solutions. The behavior of nanoaggregates is fairly complex and depends on the concentration and aging of the
solutions. The particles are stabilized in the solution by counterions. The formation of larger clusters upon dilution
of bis(1,2-dicarbollide)cobaltate(1-) solutions was observed. The secondary aggregation can be suppressed by addition
of NaCl. Gel permeation chromatography measurements of sodium bis(1,2-dicarbollide)cobaltate(1-) show that the
majority of matallacarborane molecules form nanoaggregates and only a small amount of the metallacarborane remains
molecularly soluble or forms small oligomers.

Introduction truly “soluble” in water. This fanciful group of inorganic


compounds was discovered by Hawthorne in the 1960s,9-11 and
A large group of pharmacologically active compounds could
their synthetic chemistry has been explored extensively.12,13
be characterized by a common atributeseither their hydrophobic
Metallacarboranes are based on a three-dimensional skeleton
or their amphiphilic nature. As a result, these compounds tend
built up from triangular B-H and C-H facets and a central
to form different assemblies (aggregates) in aqueous solutions1,2
metal atom(s) that is (are) sandwiched by polyhedral subclusters.
which could influence significantly their biological activity. The
They are characterized by exceptional hydrophobicity, delocalized
incorporation of a drug into an aggregate affects its physico-
negative charge, rigid geometry, low nucleophilicity, strong
chemical properties, pharmacokinetics, and interaction with
acidity of their conjugated acids, and remarkable thermal and
biological membranes.3 Moreover, in contrast to single molecules,
chemical stability.12,14
the aggregates of several small compounds were shown as unique
The above-mentioned physicochemical properties are reflected
nonspecific and promiscuous inhibitors of enzymes.4-6 The
in the ion-pairing behavior of metallacarboranes, which in turn
biological activity was determined in this case mainly by the size
led to their known applications in extraction chemistry.14-16 In
of the aggregate and its hydrophobicity. Such nonspecifity is a
several works concerning ion exchange reactions, the surface
key problem in high-throughput screening of large libraries of
activity of the [1]- anion17,18 (see Chart 1) and the hexachlorinated
chemical compounds in early-stage drug discovery.7
derivative of the [1]- anion19 at the interface of two immiscible
The recent finding of strong anti-HIV activity of polyhedral
liquids was reported. Thus, the tendency of metallacarboranes
metallacarboranes8 led us to question if these potential drugs are
to associate in aqueous solutions could be easily expected, even
when they lack an amphiphilic topology. We can learn from the
* To whom correspondence should be addressed. Phone: +420221951300.
Fax: +420224919752. E-mail: matej@vivien.natur.cuni.cz.
solution properties of a closely related group of carboranes without
† Charles University in Prague. a coordinated metal ion, 1,2-dicarba-closo-dodecaboranes de-
‡ Institute of Chemical Technology.
§ Academy of Science of the Czech Republic.
(9) Francis, J. N.; Hawthorne, M. F. J. Am. Chem. Soc. 1968, 90, 1663-1664.
(1) Atwood, D.; Florence, A. T. Surfactant Systems: Their Chemistry, Pharmacy (10) Hawthorne, M. F.; Francis, J. N. Inorg. Chem. 1971, 10, 863-864.
and Biology; Chapman and Hall: New York, 1983; Chapter 4. (11) Hawthorne, M. F.; Young, D. C.; Wegner, P. A. J. Am. Chem. Soc. 1965,
(2) Attwood, D. AdV. Colloid Interface Sci. 1995, 55, 271-303. 87, 1818-1819.
(3) Schreier, S.; Malheiros, S. V. P.; de Paula, E. Biochim. Biophys. Acta 2000, (12) Sivaev, I. B.; Bregadze, V. I. Collect. Czech. Chem. Commun. 1999, 64,
1508, 210-234. 783-805.
(4) McGovern, S. L.; Helfand, B. T.; Feng, B.; Shoichet, B. K. J. Med. Chem. (13) Bregadze, V. I.; Timofeev, S. V.; Sivaev, I. B.; Lobanova, I. A. Russ.
2003, 46, 4265-4272. Chem. ReV. 2004, 73, 433-453.
(5) Seidler, J.; McGovern, S. L.; Doman, T. N.; Shoichet, B. K. J. Med. Chem. (14) Plesek, J. Chem. ReV. (Washington, D.C.) 1992, 92, 269-278.
2003, 46, 4477-4486. (15) Rais, J.; Gruner, B. In SolVent extraction; Marcus, I., SenGupta, A. K.,
(6) McGovern, S. L.; Caselli, E.; Grigorieff, N.; Shoichet, B. K. J. Med. Chem. Eds.; Marcel Dekker: New York, 2005; p 243.
2002, 45, 1712-1722. (16) Makrlik, E.; Rais, J.; Base, K.; Plesek, J.; Vanura, P. J. Radioanal. Nucl.
(7) Feng, B. Y.; Shelat, A.; Doman, T. N.; Guy, R. K.; Shoichet, B. K. Nat. Chem. 1995, 198, 359-365.
Chem. Biol. 2005, 1, 146-148. (17) Popov, A.; Borisova, T. J. Colloid Interface Sci. 2001, 236, 20-27.
(8) Cigler, P.; Kozisek, M.; Rezacova, P.; Brynda, J.; Otwinowski, Z.; Pokorna, (18) Davion Van Mau, N.; Issaurat, B.; Amblard, G. J. Colloid Interface Sci.
J.; Plesek, J.; Gruener, B.; Doleckova, L.; Masa, M.; Sedlacek, J.; Bodem, J.; 1984, 101, 1-9.
Kraeusslich, H.-G.; Kral, V.; Konvalinka, J. Proc. Natl. Acad. Sci. U.S.A. 2005, (19) Chaumont, A.; Galand, N.; Schurhammer, R.; Vayssiere, P.; Wipff, G.
102, 15394-15399. Russ. Chem. Bull. 2004, 53, 1459-1465.

10.1021/la052201s CCC: $33.50 © 2006 American Chemical Society


Published on Web 12/16/2005
576 Langmuir, Vol. 22, No. 2, 2006 Matějı́ček et al.

Chart 1. Structures of Investigated Sandwich-Type ether. The organic layer was shaken with a 10% aqueous solution
Polyhedral Metallacarboranes Characterized by Peculiar of sodium carbonate, reextracted by water, and dried using anhydrous
Peanut-like ([1]-) and Banana-like ([2]2-) Shapes sodium sulfate. The solvent was evaporated, and solid residues were
dried in a vacuum. The sodium salts of metallacarboranes were
obtained in nearly quantitative yields.
All metallacarborane solutions were prepared by dilution of a
stock solution (c ) 10 g/L for Na[1] and Na2[2] and c ) 0.25 g/L
for Cs[1]) with deionized and filtered water.
Techniques. NMR Spectroscopy and Mass Spectrometry. 1H
(160.380 MHz) and 11B (500 MHz) NMR spectroscopy measurements
were performed on a Varian Unity-500 instrument equipped with
the appropriate decoupling accessories. All NMR spectra were
recorded from CD3COCD3 solutions at 25 °C. Chemical shift values
for 11B NMR spectra were referenced to external F3B‚OEt2, and
those for 1H{11B} NMR spectra were referenced to Si(CH3)4.
Chemical shifts δ are reported in units of parts per million downfield
from the reference. All coupling constants are reported in hertz. The
rivatized with two glucosyl units, that they behave as amphiphilic
NMR data are presented in the following format: (11B NMR) 11B
compounds, forming aggregates in aqueous solution.20 The chemical shifts δ(11B) (ppm), J(11B-1H) in parentheses; (1H{11B}
aggregation behavior was reported for other hydrophobic NMR) selectively decoupled δ(1H)-{11B selective} chemical shifts.
spherically shaped molecules, e.g., fullerene derivatives;21,22 Mass spectrometry measurements were performed on a Bruker
however, no studies were done on negatively charged polyhedral Esquire 3000 instrument using electrospray ionization. Negative
metallacarboranes. ions were detected. Samples dissolved in acetonitrile (concentrations
We have focused our research on the formation and behavior approximately 100 ng/mL) were introduced to the ion source by
of metallacarborane colloidal particles in an aqueous solution of infusion of 0.25 mL/h, the drying temperature was 300 °C, the drying
two “very water soluble” metallacarborane inhibitors of HIV gas flow 4 L min-1, and the nebulizing gas pressure 8 psi.
protease, bis(1,2-dicarbollide)cobaltate(1-) and bis[(3)-1,2- Dynamic Light Scattering (DLS). The light scattering setup (ALV,
dicarbollylcobalt]-(3,6)-1,2-dicarbacanastide(2-) ions, [1]- and Langen, Germany) consisted of a 633 nm He-Ne laser, an ALV
CGS/8F goniometer, an ALV high QE APD detector, and an ALV
[2]2-, respectively (see Chart 1), which exhibit surface activity
5000/EPP multibit, multi-τ autocorrelator. The solutions of the highest
as mentioned above. We investigate their size, shape, and behavior concentration were filtered through 0.45 µm Acrodisc filters, and
by a combination of light scattering and microscopy techniques those of lower concentrations were prepared by mixing with water
complemented by conductometry and gel permeation chroma- filtered through 0.20 µm Acrodisc filters. The measurements were
tography. To our best knowledge, this is the first report on the carried out for different concentrations and different angles at 20 °C.
molecular assembly of polyhedral metallacarboranes in solution. The data analysis was performed by fitting the measured normalized
The presented study is a part of a more complex project concerning autocorrelation function g2(t) ) 1 + β|g1(t)|2, where g1(t) is the
the physical-chemical behavior of metallacarboranes in aqueous electrical field correlation function, t is the lag time, and β is a factor
solution and their interactions with some important proteins such accounting for deviation from the ideal correlation. An inverse
as HIV protease or serum albumins (papers in preparation and Laplace transform of g1(t) with the aid of a constrained regularization
algorithm (CONTIN) provides the distribution of relaxation times,
already published8).
τA(τ):24
Experimental Section
Chemicals and Solutions. Cesium salts of metallacarboranes g1(t) ) ∫τA(τ) exp(-t/τ) d ln τ (1)
were a kind gift of Dr. Bohumı́r Grüner and Dr. Jaromı́r Plešek
(Institute of Inorganic Chemistry, Academy of Science of the Czech Diffusion coefficients were calculated from individual diffusion
Republic, Řež near Prague). They were characterized using mass modes as D ) Γ/q2, where Γ ) 1/τ and q ) (4πn0/λ) sin(θ/2) is the
spectrometry and 1H and 11B NMR spectroscopy. magnitude of the scattering vector. Here θ is the scattering angle,
Cs[1] (MW 456.65): MS m/z ) 323.7 (100) [M]- (calcd 323.740); n0 the refractive index of the pure solvent, and λ the wavelength of
1H{11B} NMR (acetone-d ) δ 3.93 (CH1,2), 3.37 (H8), 2.97 (H10),
6 the incident light. Hydrodynamic radii, RH, were evaluated from the
2.69 (H4,7), 1.92 (H9,12), 1.63 (H6), 1.56 (H5,11); 11B NMR effective diffusion coefficients using the Stokes-Einstein formula
(acetone-d6) δ 6.4 (d, B8) [142.4], 1.2 (d, B10) [138.5], -5.5 (d, RHeff ) kT/(6pDeffh), where for Deff we can write25
B9,12) [overlap], -6.2 (d, B4,7) [overlap], -17.4 (d, B5,11) [154.0],
-22.9 (d, B6) [165.5].
Cs2[2] (MW 769.08): MS m/z ) 251.6 (100) [M]2- (calcd Deff ) D(1 + CRg2q2)(1 + kDc + ...) (2)
251.635); 1H{11B} NMR (acetone-d6) δ (for atom numbering, see
Chart 1): 4.56 (CH1,2), 4.45 (H8,10), 3.42 (CH1′,2′,1′′,2′′), 3.10 where D is the angle- and concentration-independent diffusion
(H4,5,7,11), 3.06 (H8′,8′′), 2.66 (H10′,10′′), 2.42, 1.74, 1.69 (H9,- coefficient, Rg is the radius of gyration, c is the sample concentration,
12, 4′,7′,9′,12′,4′′,7′′,9′′,12′′) (overlap), 1.52 (H6′,6′′), 1.41 (H5′,- C is a constant depending on the polydispersity and the slowest
11′,5′′,11′′); 11B NMR (acetone-d6) δ 21.4 (d, B8,10) [143.5], 4.3 internal mode of motion in the scattering particle, and kD is the
(d, B8′,8′′) [143.4], 0.0 (d, B4,5,7,11) [149.5], -3.6 (d, B10′,10′′) diffusion virial coefficient. Therefore, to obtain a true RH, the apparent
[140.4], -7.9 (B9,12, 4′,7′,9′,12′,4′′,7′′,9′′,12′′) [overlap], -19.4 hydrodynamic radius has to be extrapolated to the zero scattering
(d, B5′,11′,5′′,11′′) [155.7], -24.2 (d, B6′,6′′) [167.9]. angle.
Cesium salts were converted to sodium salts using the extraction Static Light Scattering (SLS). The measurements were performed
procedure described in ref 23. The aqueous solution of the cesium on the same instrument, which was employed for DLS measurements.
salt acidified by sulfuric acid was extracted two times by diethyl
(23) Plesek, J.; Base, K.; Mares, F.; Hanousek, F.; Stibr, B.; Hermanek, S.
(20) Morandi, S.; Ristori, S.; Berti, D.; Panza, L.; Becciolini, A.; Martini, G. Collect. Czech. Chem. Commun. 1984, 49, 2776-2789.
Biochim. Biophys. Acta 2004, 1664, 53-63. (24) Berne, B. J.; Pecora, R. Dynamic Ligtht Scattering; J. Wiley-Interscience:
(21) Nakamura, E.; Isobe, H. Acc. Chem. Res. 2003, 36, 807-815. New York, 1976.
(22) Tsao, C. S.; Lin, T. L.; Jeng, U. S. J. Phys. Chem. Solids 1999, 60, (25) Burchard, W.; Schmidt, M.; Stockmayer, W. H. Macromolecules 1980,
1351-1353. 13, 1265-1272.
Molecular Assembly of Metallacarboranes in Water Langmuir, Vol. 22, No. 2, 2006 577

Data were treated by the standard method using the Zimm equation26

Kc
)
1 1
Rcor(q,c) p(q,c) Mw (
+ 2A2c + ... ) (3)

where K ) 4π2n2(dn/dc)2/λ4NA is a constant containing the refractive


index n of the solvent, refractive index increment of the colloidal
particles with respect to the solvent, dn/dc, wavelength of the incident
light, λ, and Avogadro constant, NA. The other symbols are defined
as follows: Rcor(q,c) is the corrected Rayleigh ratio, which depends
on the sample concentration and on the scattering vector q, Mw is
the weight average molar mass of scattering particles, and A2 is the
“light-scattering-weighted” second virial coefficient of the concen-
tration expansion.
The function p(q,c) ) Icor(q)/Icor(q ) 0) takes into account both
intraparticle and interparticle interference effects. We assume that Figure 1. DLS correlation spectra of Na[1] aggregates 2 days after
the scattering function p(q,c) may be reasonably expressed as a preparation for different concentrations: (curve 1) 0.05 g/L, (curve
product of the single particle form factor P(q) and the solution 2) 0.3 g/L, (curve 3) 0.5 g/L, and (curve 4) 5 g/L.
structure factor S(q,c), i.e., p(q,c) ) P(q) S(q,c).27 The structure
factor was equal to 1 for all concentrations and samples studied out at the laboratory temperature (22 ( 2 °C). A mixture of standards,
herein. Since the values of the refractive index increment dn/dc and MIX A (consisting of glucose and three dextrans), was used for
the actual concentration of scattering particles are unknown in our relative calibration.
system, SLS data were used only to obtain the form factor P(q) by
dividing extrapolated values of Kc/R(c) to the zero angle by apparent Results and Discussion
values of Kc/R(q,c), where c is concentration of metallacarboranes Characterization of Metallacarborane Solutions by Light
in solution and the dn/dc value is arbitrary. From a Guinier region Scattering. The solution behavior and aggregation of metal-
of the form factor the radius of gyration, Rg, was calculated. lacarborane were studied by static and dynamic light scattering.
All solutions of metallacarboranes, measured by light scattering, DLS measurement reveals the presence of relatively large colloidal
were colored (yellow for [1]- or deep red for [2]2-); however, the
wavelength of the used laser (633 nm) does not coincide with
particles containing a large number of aggregated metallacar-
absorption bands of metallacarboranes in water. Therefore, all SLS borane molecules in all studied aqueous solutions of Na[1], Na2-
and DLS results are reliable.26 [2], and Cs[1]. The solutions are stable and do not precipitate.
Atomic Force Microscopy (AFM). All measurements were In this paper we focused mainly on the behavior of Na[1] in
performed in the tapping mode under ambient conditions using a aqueous solutions; however, the behavior of Na2[2] and Cs[1]
commercial scanning probe microscope, Digital Instruments Nano- is discussed for comparison to show a more general property of
Scope dimensions 3, equipped with a Nanosensors silicon cantilever, metallacarborane aggregates in water.
typical spring constant 40 N m-1. Metallacarborane particles were Distributions of hydrodynamic radii RH of aggregates of Na-
deposited on a fresh (i.e., freshly peeled out) mica surface (flogopite, [1] in aqueous solution, measured 2 days after the preparation,
theoretical formula KMg3AlSi3O10(OH)2, Geological Collection of are plotted in Figure 1 for several concentrations. The position
Charles University in Prague, Czech Republic) by a fast dip coating
in an aqueous solution of sample. After the evaporation of water,
and width of the peaks obtained from the DLS autocorrelation
the samples for AFM were dried in a vacuum oven at ambient function vary with concentration. The aggregates are fairly
temperature for ca. 5 h. monodisperse at low concentrations (curve 1). Smaller aggregated
Scanning Electron Microscopy (SEM). SEM measurements were particles form at medium concentrations (curves 2 and 3), and
carried out using a scanning electron microscope coupled with an the fraction of large particles increases substantially in the region
energy-dispersive X-ray analyzer (Hitachi S-4700 + Noran D-6823). of high concentrations (curve 4). Our study showed that this
Metallacarborane particles were deposited on glass slides by a fast trend is very reproducible. The particle polydispersity increases
dip coating in an aqueous solution of sample. After the evaporation with concentration, but it changes also with time. The changes
of water, the slides were dried in a vacuum oven at ambient in particle size with time are depicted in Figure 3 and in the inset
temperature for ca. 5 h. The surface was than gold coated, and the in Figure 2. Values of the hydrodynamic radius RH and radius
micrographs were taken at an acceleration voltage of 10 or 15 kV.
Conductometry. The conductivity of aqueous metallacarborane
of gyration Rg of aggregates 1 month after filtration of the solutions
solutions was measured with precision conductivity equipment are shown in Figure 2 as functions of Na[1] concentration. In
consisting of a Radiometer apparatus CDM210 and a conductivity the inset, hydrodynamic radii RH of several concentrations of
cell. The conductivity was measured for solutions of different Na[1] measured at different times after preparation are shown.
concentrations, which were prepared by dilution of a stock solution In the concentration region above 1 g/L, the curves increase with
(1 g/L) with deionized water. All measurements were carried out increasing concentration and the polydispersity of the aggregates
at a temperature of 25 °C. is very high immediately after preparation. The polydispersity
Gel Permeation Chromatography (GPC). The GPC equipment of particles in the solution remains high, even if the solution has
(Watrex, Prague, Czech Republic) consisted of a Delta Chrom SDS been filtered and a certain fraction of the sample has been found
030 pump, a Delta Chrom UVD 200 detector, and a Rheodyne to adsorb on the filter. This fact indicates that large aggregates
injection valve, model 7125 (Cotati, CA) with a 20 µL sample loop.
Signal acquisition and data handling were performed with the
or clusters form in the solution immediately after filtration. The
DataApex Clarity 2.1 software. A commercially available steel radius of particles decreases during aging of the solutions (several
column, Watrex GMB 10000 (column length 250 mm), particle size weeks after the preparation) and finally remains constant in the
7 µm, and deionized water as the mobile phase were used. The concentration region above 0.3 g/L. This indicates that unstable
detection wavelength was 214 nm. The measurements were carried clusters, which form shortly after the preparation of solutions,
disappear, because they transform slowly into relatively mono-
(26) Kratochvil, P. In Classical Light Scattering from Polymer Solutions, disperse stable ones.
Polymer Science Library 5; Jennkins, A. D., Ed.; Elsevier: Amsterdam, 1987.
(27) Klein, R.; D’Aguanno, B. In Light Scattering. DeVelopments and Principles; The behavior of very dilute solutions of Na[1] is surprising.
Brown, W., Ed.; Clarendon Press: Oxford, England, 1996; p 30. The aggregates are fairly monodisperse, and their radius increases
578 Langmuir, Vol. 22, No. 2, 2006 Matějı́ček et al.

Figure 4. Inverse form factor P-1 of Na[1] aggregates as a function


Figure 2. Hydrodynamic radius RH (curve 1) and radius of gyration of q2Rg2 1 h (curve 1), 3 h (curve 2), 1 day (curve 3), 2 days (curve
Rg (curve2) as a function of the Na[1] concentration 1 month after 4), and 1 week (curve 5) after preparation. The Na[1] concentration
preparation. Inset: RH of Na[1] aggregates as a function of the is 0.05 g/L. The solid curves organize the data visually but do not
concentration 1 day (curve 1, hollow spheres), 1 week (curve 2, represent an attempt to model the data.
black spheres), 2 weeks (curve 3, hollow squares), and 1 month
(curve 4, black squares) after preparation. The solid lines connecting of aggregates in Na2[2] solutions does not change with
the data are drawn only to guide the eye. concentration and yields an RH of ca. 110 nm. This metallac-
arborane is fairly soluble in water. It contains two Na+ ions in
each molecule, and it seems that Na2[2] aggregates are sufficiently
stable in water and therefore do not form secondary clusters.
Information on the shape of the aggregates can be obtained
from the form factor, P(q), in combination with the ratio Rg/RH.
The static light scattering measurements did not show any
evidence of interparticle interactions, which are frequently
observed in charged colloidal solutions at low ionic strength.
Hence, the structure factor is expected to equal 1, and only the
form factor (intraparticle effects) influences the light scattering
function p(q,c) (see eq 3). The ratios Rg/RH ) 0.785 ( 0.112,
0.736 ( 0.060, and 0.704 ( 0.112 for Na[1], Cs[1], and Na2[2],
respectively, have been obtained from SLS and DLS measure-
Figure 3. Hydrodynamic radius RH (curve 1) and radius of gyration ments in aqueous solutions. For monodisperse hard spheres the
Rg (curve 2) of Na[1] aggregates as a function of time. The Na[1] Rg/RH ratio equals 0.775, for hard ellipsoids this number goes
concentration is 0.05 g/L. The solid lines connecting the data are sharply up, and for hollow spheres the value is about 1.28 The
drawn only to guide the eye. measured values suggest that metallacarborane solutions contain
roughly spherical particles.
with time (Figure 3). The increase at early times is rather fast, An additional proof of this assumption is the shape of the
and then it slows, but we found that it continues at least 1 month inverse form factor, P-1(q), as a function of the product Rg2q2
after preparation. The system seems to reach equilibrium (i.e., (Figure 4).26 The function turns sharply up and exhibits a
the radius of the aggregates does not change) after several months. maximum at Rg2q2 ) ca. 12, which is predicted by theory for
The size of the aggregates increases with decreasing concentration homogeneous spheres. The shape of the function P-1(q) can be
below 0.3 g/L. It is worth mentioning that all time evolutions influenced by various factors, e.g., by polydispersity in sizes,
were found to be reproducible. and does not always give unambiguous information on the particle
The behavior of metallacarboranes is fairly complex, and its shape. However, the observed shape of the function P-1(q) with
understanding requires knowledge of the shape of the particles a clearly pronounced maximum provides a fair indication of a
and their fraction in the solution. Because the intensity of scattered more or less spherical shape. Curves measured at different times
light is very sensitive to the presence of large aggregates, the (Figure 4) demonstrate the growth of spherical particles with
molecularly soluble metallacarboranes and their oligomers do time. For relatively small particles existing at early times (curves
not almost contribute to the light scattering signal, and the 1-4), the product Rg2q2 does not reach enough large values to
estimation of the fraction of aggregates in the solution cannot show a maximum and the maximum is detectable only for
be obtained from SLS and DLS data. This is why we have to solutions more than 1 week old (curve 5). To yield convincing
use a combination of several other methods (e.g., GPC; see later). proof for the spherical shape of metallacarborane particles, we
The behavior of Cs[1] was studied for comparison. It was investigated the system also by other techniques (e.g., SEM and
found that the solubility of Cs[1] is low (about 0.25 g/L) and the AFM; see later).
sizes of the aggregates increase strongly with dilution. The Assuming the spherical shape of aggregates, the number of
concentration dependence of RH is similar to that of Na[1] (not aggregated metallacarborane molecules can be estimated from
shown). Furthermore, all distributions of DLS correlation times the particle diameter (measured by DLS) and density (obtained
for Cs[1] solutions were monomodal and fairly narrow in the from the literature29,30). Unfortunately, there are no data available
whole concentration range.
It is worth mentioning that no steep increase in particle size (28) Burchard, W. In Light Scattering. DeVelopments and Principles; Brown,
W., Ed.; Clarendon Press: Oxford, England, 1996; p 439.
with decreasing metallacarborane concentration was observed (29) Zalkin, A.; Hopkins, T. E.; Templeton, D. H. Inorg. Chem. 1967, 6,
in solutions of Na2[2]. Consequently, the hydrodynamic radius 1911-1915.
Molecular Assembly of Metallacarboranes in Water Langmuir, Vol. 22, No. 2, 2006 579

example, by Langmuir in 193832). The classical DLVO theory


(Derjaguin-Landau-Verwey-Overbeek)33,34 assumes only re-
pulsive electrostatic interactions between two equally charged
particles. To explain the experimental results, which cannot be
explained by the DLVO model, another pairwise potential was
proposed, which contains a long-range attractive part.35 The
tendency to form clusters or concentrated domains of ordered
particles and dilute regions was observed in solutions of silica
or latex spheres.36-39 The destabilization of a mixture of colloidal
particles differing in size was observed and explained by depletion
forces and phase separation at low ionic strength.40,41 Furthermore,
many experimental studies performed over several past decades
Figure 5. Dependence of relaxation rates Γ obtained by DLS on pointed out the importance of counterions.42 The concept of so-
the magnitude of the scattering vector of Na[1] aggregates in water called “thermodynamically bound” counterions explains the phase
with concentrations of 10 g/L (curve 1), 0.27 g/L (curve 2), and 0.07 transition at low ionic strength43 and the formation of temporal
g/L (curve 3). Linear dependence is assumed in the case of a diffusive aggregates44 in colloidal systems. In dilute solutions, macroion
motion of particles.
clusters stabilized by counterions have usually been observed.45
The behavior cannot be explained by a simple pairwise potential.
for Na[1], but the density of this compound can be expected to It is necessary to consider the net electrostatic interaction of the
be in the range 1-2 g/cm3. It should be kept in mind that the assembly of large colloidal particles and small ions.42
density of the aggregates could be lower than that in the solid
The above conclusions can be supported by the following
state because of hydration and repulsion of anions (counterions
experiment concerning the destabilization of metallacarborane
partially release toward the bulk solution). Therefore, only a
aggregates upon dilution. If Na[1] solution is diluted by 0.1 M
rough estimate of the number of molecules in each particle,
NaCl instead of pure water, no cluster formation occurs in the
106-107, was done.
low concentration region in salted solution and the hydrodynamic
To ensure that the measured hydrodynamic radii are reliable,
radius of the aggregates remains the same as that in the high
we have to prove that the motion of metallacarborane particles
concentration region in pure water (ca. 110 nm). Hence, the
in aqueous solutions corresponds to diffusion.28 The proof of the
addition of salt suppresses the formation of secondary aggregates
diffusion-controlled motion is the linear dependence of the
in dilute solutions as predicted by theory.
relaxation rate Γ on q2, i.e., Γ ) Dq2, where D is the diffusion
coefficient. This dependence is depicted in Figure 5 for three The conductivity measurements on dilute Na[1] solutions were
different concentrations of Na[1] in water. The agreement is carried out to obtain the critical micellar concentration, cmc.
reasonably good within the range of experimental errors; The specific conductivity of associating systems as a function
nevertheless, the slope of the log(Γ)-log(q) dependence, which of concentration shows typically one or two breaks corresponding
should be 2, is slightly higher and gradually increases with q (up to the critical micellar and premicellar concentrations.2,46 We
to ca. 2.25) especially for clusters below the concentration limit did not observe any breaks up to 2 g/L (conductometer limit),
at 0.3 g/L. The experimental slope may be slightly higher than despite the existence of aggregates in Na[1] solutions. The
2, because it contains a contribution of a slow internal mode dependence of specific conductivity on concentration is almost
proportional to q3. Its origin can be attributed to the motion of linear (not shown). It confirms our conclusion from light scattering
primary aggregates in the secondary clusters, mainly in solutions measurements that aggregates exist in the whole measured
below the limit of 0.3 g/L.31 concentration range. Therefore, the detection of the cmc is below
The performed light scattering measurements suggest that the limits of both light scattering and conductivity measurements.
aqueous solutions of metallacarborane contain aggregates of We have to keep in mind that the detection of the cmc is
roughly spherical shape. The behavior of the aggregates is complicated by the observed formation of large clusters in very
complex and depends on the concentration and aging of the dilute solutions. It can be suppressed by the addition of salt, but
solution. Large clusters of primary aggregates are formed, when this precludes conductometry measurements. The surface tension
solutions of Na[1] and Cs[1] are diluted below ca. 0.3 g/L. This measurements of very dilute solutions might be more appropriate
observation may be rationalized as follows. Individual aggregates even after addition of the salt as pointed out by a reviewer of
are stabilized by counterions. After dilution, the concentration this paper. These experiments have not yet been carried out.
of counterions near the surface of metallacarborane particles
decreases, because they escape in bulk solution and gain (32) Langmuir, I. J. Chem. Phys. 1938, 6, 873-896.
(33) Deryagin, B.; Landau, L. Acta Physicochim. URSS 1941, 14, 633-662.
translational entropy. Consequently, the system is destabilized (34) Verwey, E. J. W.; Overbeek, J. Th. G. Theory of the Stability of Lyophobic
and starts to aggregate to decrease the overall surface of the Colloids; Elsevier Publishing Co., Inc.: New York, 1948.
(35) Sogami, I.; Ise, N. J. Chem. Phys. 1984, 81, 6320.
aggregates. This means that the clustering of aggregates is a (36) Ito, K.; Ieki, T.; Ise, N. Langmuir 1992, 8, 2952-2956.
result of the entropy-to-enthalpy interplay. Hydrophobic parts of (37) Yoshida, H.; Ise, N.; Hashimoto, T. J. Chem. Phys. 1995, 103, 10146-
metallacarborane molecules try to avoid the polar aqueous 10151.
(38) Yoshida, H.; Ise, N.; Hashimoto, T. Langmuir 1995, 11, 2853-2855.
medium, and the primary aggregates form large clusters at low (39) Dosho, S.; Ise, N.; Ito, K.; Iwai, S.; Kitano, H.; Matsuoka, H.; Nakamura,
dilution when small ions do not condense at their surface and H.; Okumura, H.; Ono, T. Langmuir 1993, 9, 394-411.
do not provide sufficient stability. (40) Henderson, D.; Boda, D.; Chan, K. Y.; Wasan, D. T. Mol. Phys. 1998,
95, 131-135.
The assumption that equally charged colloidal particles (41) Xu, W.; Nikolov, A. D.; Wasan, D. T. AIChE J. 1997, 43, 3215-3222.
surrounded by small ions can effectively interact via attractive (42) Schmitz, K. S. Langmuir 1997, 13, 5849-5863.
(43) Lin, S. C.; Lee, W. I.; Schurr, J. M. Biopolymers 1978, 17, 1041-1064.
forces has been discussed in the literature since the 1930s (for (44) Clark, N. A.; Ackerson, B. J.; Hurd, A. J. Phys. ReV. Lett. 1983, 50,
1459-1462.
(30) St. Clair, D.; Zalkin, A.; Templeton, D. H. Inorg. Chem. 1969, 8, 2080- (45) Groehn, F.; Antonietti, M. Macromolecules 2000, 33, 5938-5949.
2086. (46) Hunter, R. J. Foundations of Colloid Science; Clarendon Press: Oxford,
(31) Wu, C.; Ngai, T. Polymer 2004, 45, 1739-1742. 1995; Vol. I.
580 Langmuir, Vol. 22, No. 2, 2006 Matějı́ček et al.

Characterization of Metallacarborane Aggregates by


AFM and SEM. Light scattering measurements revealed the
existence of aggregates in metallacarborane solutions, but a
detailed study of their structure requires more direct experimental
techniques, e.g., AFM and SEM. Although the shape of the
colloidal particles is often deformed after their deposition on the
substrate, the microscopy study yields valuable information on
the shape and polydispersity of aggregates and also on the structure
of detected clusters. By preparing samples for microscopy
measurements from dilute solutions, we ensure that particles are
evenly spread and do not overlap on the substrate. Hence, we
can relate the original shape of the particles in solution and that
on the surface.
The AFM scan of Na2[2] aggregates on a mica surface (Figure
6a) shows fairly monodisperese particles with a spherical plan
of view deposited from the solution with a concentration of 0.5
g/L. The aggregates of Na[1] deposited from solutions in the
concentration range 0.25-10 g/L are very similar. On the contrary,
the aggregates of Na[1] deposited on a mica surface from very
dilute solutions (0.03 g/L) are larger and polydisperse (Figure
6b). In Figure 6c, typical profiles of Na[1] aggregates for different
concentrations are shown. The dependence of aggregate sizes on
concentration is similar to that in the solutions (compare Figure
2). Nevertheless, the clusters deposited on the surface are less
uniform than those in the solution. By scanning a large mica
surface, we find both regions of small spherical aggregates and
their clusters. This observation supports the assumption concern-
ing the phase separation in very dilute solutions of metallac-
arboranes. It is necessary to take into account that large clusters
on mica may split into smaller fragments due to deformation
forces acting at the surface.
The SEM pictures of the gold-coated Na[1] sample deposited
on glass from the solution with a concentration of 0.175 g/L also
show (Figure 7a,b) both individual spherical aggregates and
clusters formed by smaller particles for the same sample. This
observation supports the suggested mechanism of secondary
aggregation (clustering) below the concentration limit of 0.3
g/L. It should be mentioned that the coexistence of small
aggregates and their clusters was observed for all concentrations
of Na[1] solutions aged several days (0.03, 0.175, 2.0, and 10.0
g/L). This indicates that clusters form in the whole concentration
range with time. However, they are less stable in the concentration
region above ca. 0.3 g/L (see the light scattering section).
Both microscopy techniques image metallacarborane ag-
gregates as relatively spherical and monodisperse particles.
Experiments with samples deposited from solutions differing in
aging time show that metallacarboranes undergo secondary
aggregation with time and dilution.
Fraction of Na[1] Aggregates in Aqueous Solutions. To
obtain information on the fraction of aggregated metallacarborane
molecules, a gel permeation chromatography study of Na[1]
solutions was carried out. All chromatograms for concentrations
between 1 and 10 g/L are fully reproducible and show two peaks
at retention times of about 6.5 and 13.0 min (Figure 8). For
solutions with c < 1 g/L, the second peak is hidden in the noise
and cannot be analyzed. The first peak can be assigned to
aggregates and the second one to the fraction of Na[1] molecules Figure 6. (a) AFM scan on a mica surface of Na2[2] aggregates
or small oligomers (dextran calibration), which are detectable adsorbed from solution with a concentration of 0.5 g/L. (b) AFM
neither by light scattering nor by microscopy. scan on a mica surface of Na[1] aggregates adsorbed from solution
with a concentration of 0.03 g/L. (c) Typical profiles of AFM scans
Because the kinetics of aggregate formation is relatively slow of Na[1] aggregates deposited on a mica surface from solution with
(see the light scattering section), the ratio of large aggregates to concentrations of (curve 1) 0.03 g/L, (curve 2) 0.1 g/L, and (curve
small oligomers should remain unchanged during the separation 3) 0.25 g/L.
Molecular Assembly of Metallacarboranes in Water Langmuir, Vol. 22, No. 2, 2006 581

stationary phase. Hence, the equilibrium between molecules and


aggregates of metallacarboranes may shift in the column, and
the amount of aggregates in solution can be underestimated by
GPC. The fraction of Na[1] in the aggregates was calculated
from the measured chromatograms as the ratio of the first peak
area divided by the area of both peaks. Its value does not change
with concentration, and it yields 0.82. This means that the majority
of Na[1] molecules (at least 80%) form in relatively large
aggregates.
Conclusions
The metallacarboranes studied in this work show a pronounced
amphiphilic character, which results in formation of aggregates
in aqueous solutions. Their size, shape, and behavior were studied
by a combination of light scattering (SLS and DLS) and
microscopy (AFM and SEM) techniques with the help of
chromatographic measurements. The association of metallac-
arboranes in water was observed for the first time.
Several weeks after the preparation, the hydrodynamic radii
of the aggregates reach “equilibrium” values around 115 nm for
Na[1] above the concentration limit of 0.3 g/L and 110 nm for
Na2[2] in the whole concentration range studied. Angular SLS
measurements, together with values of the Rg/RH ratio and
microscopy scans, suggest that the aggregates are spherical.
Individual particles of Na[1] and Cs[1] undergo secondary
aggregation when the solution is diluted below 0.3 g/L. The size
of the clusters formed at high dilution increases steeply with
decreasing concentration. The clustering is efficiently suppressed
by the addition of an indifferent electrolyte, e.g., NaCl.
The fraction of Na[1] molecules which form the aggregates,
determined by GPC, is approximately 80%. This means that the
majority of metallacarborane molecules form aggregates in
aqueous solutions and only a relatively small fraction of molecules
Figure 7. SEM scans of Na[1] aggregates deposited from aqueous
solution with a concentration of 0.175 g/L on a glass surface and dissolve molecularly or form small aggregates.
gold coated: (a) individual particles of aggregates and (b) clusters
of aggregates. Acknowledgment. We express our gratitude to Dr. Jaromı́r
Plešek and Dr. Bohumı́r Grüner (Institute of Inorganic Chemistry,
Academy of Science of the Czech Republic, Řež near Prague,
Czech Republic) for providing the samples of metallacarboranes
and for helpful discussion. We also thank Dr. Eva Tesařová
(Department of Physical and Macromolecular Chemistry, Charles
University in Prague, Czech Republic) for GPC measurements
and data analysis, Ing. Milena Špı́rková (Institute of Macro-
molecular Chemistry, Academy of Science of the Czech Republic
in Prague) for AFM measurements, and Dr. Jan Konvalinka and
Milan Kožı́šek (Institute of Organic Chemistry and Biochemistry,
Academy of Science of the Czech Republic, Prague, Czech
Republic) for fruitful research cooperation following this project.
This work was supported by grants from the Ministry of Education
of the Czech Republic (MSM 6046137307, and LC 512), Grants
Figure 8. Gel permeation chromatogram of a Na[1] aqueous solution 303/05/0336, 203/03/0900, 203/03/0716, and 301/04/1315 from
with a concentration of 2 g/L. the Grant Agency of the Czech Republic, and EU Grant CIDNA
NMP4-CT-2003-505669 to V.K. K.P. and P.M. acknowledge
in the column.47 Nevertheless, the mobility of molecularly soluble the financial support of the Grant Agency of the Czech Republic
metallacarborane or small oligomers (based on the “second” (Grant 203/04/0490). This work was partly supported by the
peak) is slower than the water flow. This is probably due to grant from Ministry of Education (MSMT) of Czech Republic
specific interactions of metallacarborane molecules with the within Programme 1M0508 “Research Centre for new Antivirals
and Antineoplastics” to P.C.
(47) Prochazka, K.; Bednar, B.; Tuzar, Z.; Kocirik, M. J. Liq. Chromatogr.
1988, 11, 2221-2239. LA052201S

You might also like