You are on page 1of 4

COMMUNICATIONS

[5]
[6]
[7]
[8]
[9]

A. Tracz, J. K. Jeszka, M. D. Watson, W. Pisula, K. Mllen, T. Pakula,


J. Am. Chem. Soc. 2003, 125, 1682.
F. C. Grozema, T. J. Savenije, M. J. W. Vermeulen, L. D. A. Siebbeles,
J. M. Warman, A. Meisel, D. Neher, H.-G. Nothofer, U. Scherf, Adv. Mater. 2001, 13, 1627.
A. M. van de Craats, J. M. Warman, A. Fechtenktter, J. D. Brand, M. A.
Harbison, K. Mllen, Adv. Mater. 1999, 11, 1469.
a) H. Monobe, K. Awazu, Y. Shimizu, Adv. Mater. 2000, 12, 1495.
b) H. Monobe, K. Awazu, Y. Shimizu, Mol. Cryst. Liq. Cryst. 2001, 364,
453.
J. C. Wittmann, P. Smith, Nature 1991, 352, 414.

Direct Synthesis of Se@CdSe Nanocables and


CdSe Nanotubes by Reacting Cadmium Salts with
Se Nanowires**
By Xuchuan Jiang, Brian Mayers, Thurston Herricks,
and Younan Xia*
Templating against currently existing nanowires (or rods,
belts) provides a straightforward and powerful route to
greatly expand the diversity of materials that can be processed
as uniform, one-dimensional (1D) nanostructures.[1] In one
approach, the surfaces of nanowires could be directly coated
(using a range of different methods) with conformal sheaths
made of a different material to generate coaxial nanocables.[2]
Subsequent removal of the nanowires would lead to the formation of nanotubes with well-controlled dimensions. In another approach, it has been demonstrated that single crystalline nanowires could serve as substrates for the epitaxial
growth of another material to obtain coaxial, bilayer nanotapes characterized by sharp structural and compositional interfaces.[3] By carefully modulating the composition of reactant in sequential steps, it is also possible to fabricate
semiconductor multiple-sheath nanowire heterostructures via
epitaxial growth.[4] In a third approach (or the so-called template-engaged process), nanowires had been partially (e.g., on
the surface only) or completely converted to other materials
without changing the 1D morphology when they were reacted
with appropriate chemical reagents under carefully controlled
conditions.[5] The concept of this method was originally demonstrated by Lieber and co-workers, where they found that
highly crystalline nanorods of metal carbides could be formed

[*] Prof. Y. Xia, Dr. X. Jiang, Dr. B. Mayers


Department of Chemistry
University of Washington
Seattle, WA 98195 (USA)
E-mail: xia@chem.washington.edu
T. Herricks
Department of Materials Science and Engineering
University of Washington
Seattle, WA 98195 (USA)

[**] This work has been supported in part by the STC Program of the National
Science Foundation (NSF) under Agreement Number DMR-0120967, a
Career Award from the NSF (DMR-9983893), and a Fellowship from the
David and Lucile Packard Foundation. Y. X. is a Camille Dreyfus Teacher
Scholar (2002) and an Alfred P. Sloan Research Fellow (2000). T. H. and
B. M. thank the Center for Nanotechnology at the UW for the IGERT
Fellowship Awards supported by the NSF (DGE-9987620).

1740

2003 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

by reacting carbon nanotubes with the vapors of metal oxides


or halides at elevated temperatures.[5a] A similar procedure
(including the use of both vapor- and solution-phase reactions) was later exploited by many research groups to generate 1D nanostructures from a wealth of solid materials. These
studies have also made it possible to incorporate a number of
functions (e.g., luminescent, ferromagnetic, ferroelectric,
piezoelectric, and superconducting) into an individual nanowire that will find applications in various areas. Here, we
would like to add another example to this list, where uniform
nanowires of t-Se were employed as chemical templates to
generate Se@CdSe nanocables and then CdSe nanotubes.
Figure 1 shows a schematic outline of the approach. The
first step involved the synthesis of single-crystalline nanowires
of t-Se via a sonochemical process.[6] When refluxed in an

Fig. 1. Schematic illustration describing the formation of a Se@CdSe nanocable


and then a CdSe nanotube when the t-Se nanowire was refluxed with Cd2+ in an
aqueous medium, followed by removal of the unreacted core of t-Se through
thermal evaporation.

aqueous medium containing Cd2+ cations, elemental Se disproportionated into Se2 and SeO32 anions.[7] The Se2 anions
then combined with the Cd2+ cations to generate insoluble
nanoparticles made of CdSe, which were deposited in situ as a
conformal sheath around each t-Se template to produce a
Se@CdSe nanocable structure. The major reactions involved
in this process can be summarized as the following:
2
3Se(s) + 3H2O 2Se(aq)
+ SeO32(aq) + 6H+(aq)

(1)

2
+ SeO32(aq) + 2Cd2+
Se(aq)
(aq) CdSe(s) + CdSeO3 (s)

(2)

Note that the solubility of CdSeO3 is sufficiently high at the


refluxing temperature that it would not inhibit the formation
of a dense CdSe sheath around each t-Se template via processes such as co-precipitation with CdSe. As the reaction was
cooled down to room temperature, CdSeO3 did precipitate
out as nanoparticles decorating the surfaces of Se@CdSe
nanocables. Fortunately, this unwanted byproduct could be effectively removed by washing the as-synthesized sample with
hot water in the setting of filtration. Due to the relatively low
melting point of t-Se (~ 217 C) as compared with CdSe nanoparticles,[8] the unreacted core of t-Se could be conveniently
removed through evaporation (by heating at ~ 230 C for a
few minutes) to obtain the nanotube made of pure CdSe.
DOI: 10.1002/adma.200305737

Adv. Mater. 2003, 15, No. 20, October 16

**

20

(110)

30

40

(210)

(211)

(202)

(201)

(112)

(C)

10

CdSeO3

(103)

**

(112)

(110)
(102)
(111)

(100)

(B)

(100)
(002)
(101)

Intensity (a.u.)

(A)

(D)

* CdSe

(101)

Se

50

60

70

2 (degrees)
Fig. 2. XRD patterns showing materials that were involved at different stages
of the reaction: A) single crystalline nanowires of t-Se; B) a mixture containing
Se@CdSe nanocables and CdSeO3 solid; C) Se@CdSe nanocables; and
D) CdSe nanotubes.

Adv. Mater. 2003, 15, No. 20, October 16

http://www.advmat.de

ter) prior to reaction with the aqueous Cd2+ solution. All of


the peaks could be indexed to the trigonal phase of selenium.
Figure 2B shows the XRD pattern taken from an as-synthesized sample after the reaction had proceeded for a period of
~ 8 h. The appearance of new peaks associated with CdSe and
CdSeO3 indicates the transformation of t-Se into these two
materials as described in Equation 2. The relatively high
abundance of t-Se phase in this sample suggests that the reactions only occurred on the surfaces of the t-Se nanowires, with
Se@CdSe nanocables as the primary product (see Fig. 1). The
byproduct, CdSeO3, was completely soluble at the refluxing
temperature, and it only precipitated out as decoration particles on the surfaces of nanocables when the reaction solution
was cooled down to room temperature. For the same reason,
the CdSeO3 particles could be conveniently removed by washing the sample with hot water to generate pure Se@CdSe
nanocables (Fig. 2C). Finally, nanotubes made of CdSe (see
Fig. 2D) could be prepared by heating the sample at ~ 230 C
for a few minutes to drive off the unreacted t-Se cores. The
broadening of all diffraction peaks indicates that the CdSe
nanotubes were polycrystalline in structure, with their walls
composed of nanoparticles having relatively small dimensions.
We have also investigated this reaction under various conditions, and the one described in the Experimental section
seems to be the best.
Figure 3 shows transmission electron microscopy (TEM)
images detailing the structural and morphological changes involved in this template-engaged reaction. Figure 3A shows a
typical TEM image of the templatessingle-crystalline nanowires of t-Se that had a mean diameter around 50 nm. Figure 3B shows a TEM image of Se@CdSe nanocables after the
CdSeO3 byproduct had been selectively removed by washing
with hot water. Obviously, the surfaces of t-Se nanowires had
become rougher due to the formation of CdSe sheaths in the
form of aggregated nanoparticle. Figure 3C shows a typical
TEM image of the sample after the unreacted t-Se cores had
been removed through thermal evaporation, with the contrast
clearly depicting a tubular morphology. The inset gives a typical TEM image taken from the microtomed sample, clearly
showing the cross-sections of these nanotubes with a wall
thickness in the range of 510 nm. Figure 3D shows the TEM
image of another nanotube sample where the t-Se was broken
(via sonication) into relatively short rods before reacting with
the aqueous Cd2+ solution. In this case, cylindrical shells with
closed ends were obtained as the final product. The inset of
this figure shows an electron diffraction pattern taken from a
random assembly of such shells. The appearance of diffusive
rings indicates the formation of polycrystalline walls, with the
composition indexed to pure CdSe.
We have also measured the changes in optical properties accompanying the formation of Se@CdSe nanocables and CdSe
nanotubes. Figure 4A compares the ultraviolet-visible (UVvis) absorption spectra taken from aqueous suspensions of
t-Se nanowires, and the corresponding Se@CdSe nanocables
and CdSe nanotubes. The extinction peak was slightly redshifted (by ~ 13 nm) as the surfaces of Se nanowires were

2003 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

1741

COMMUNICATIONS

Although the transformation of 1D nanostructures into different materials with a similar or complementary morphology
is not a new concept,[5] the demonstration described here
represents the first attempted synthesis of CdSe nanotubes
through a template-engaged reaction. As one of the most important semiconductors for optoelectronics, CdSe nanostructures (mostly in the form of nanoparticles or quantum dots,[9]
and most recently in the form of nanorods and tetrapots[10])
have been extensively exploited for use in the fabrication of
light-emitting diodes,[11] lasers,[12] solar cells,[13] and biomedical fluorescent tags.[14] However, the synthesis of CdSe nanotubes has been met with limited success, and literature search
indicated that only a few groups were able to generate tubular
nanostructure of CdSe by templating against cylindrical micelles assembled from organic surfactants.[15] The lengths of
nanotubes synthesized using such an approach were often restricted to the scale below 25 lm. The difficulty in selecting
an appropriate surfactant to control the morphology also limits the extension of this method to other semiconductors. In
comparison, the approach described here seems to represent
a generic one, and can be potentially extended to most binary
chalcogenides.[16] Moreover, the dimensions of resultant nanotubes were determined by those of the t-Se nanowires, which
could be readily synthesized as monodisperse samples, in copious quantities, and with controllable diameters in the range
of 10100 nm and lengths up to 50 lm.[6] As a result, we expected that the method described here will open the door to
both fundamental studies and technological applications related to tubular nanostructures made of various chalcogenides.
We first used powder X-ray diffraction (XRD) to follow the
progression of this reaction and to identify all the phases involved in this process. Figure 2 shows the XRD patterns obtained at different stages of this synthesis. Figure 2A shows
the XRD pattern of pure t-Se nanowires (~ 50 nm in diame-

COMMUNICATIONS
Fig. 3. TEM images showing: A) single crystalline nanowires of t-Se, B) Se@CdSe nanocables, and C) CdSe nanotubes. The inset of (C) shows the cross-sections of three CdSe nanotubes that were obtained from a microtomed sample. The thickness of these CdSe nanotubes was in the range of 510 nm. D) TEM image of CdSe nanotubes that
were formed by templating against relatively short nanorods of t-Se. The inset shows a typical electron diffraction
pattern, where all rings could be indexed to the hexagonal phase of CdSe.

transformed into thin sheaths of CdSe. Once the unreacted Se


cores had been removed, the UV-vis absorption spectrum exhibited features resembling those characteristic of CdSe quantum dots, with a broad, well-resolved peak located at
~ 652 nm. Relative to the value reported for the bulk crystal
of CdSe, the absorption edge was blue-shifted by ~ 61 nm.[17]
This observation further confirms that the CdSe nanotubes
synthesized using the present procedure were polycrystalline
in structure, with an average domain size around 8 nm (as derived from the position of the extinction peak).[18] Figure 4B
compares the photoluminescence spectra recorded from
aqueous suspensions of these three different samples. While
the t-Se nanowires displayed no detectable fluorescence signal
in the visible region, both Se@CdSe nanocables and CdSe
nanotubes exhibited a strong fluorescence emission with essentially the same profile (with a peak located at ~ 667 nm).
From the emission spectra, the size of CdSe nanoparticles that
made up the polycrystalline walls of the cable or tubular structures was also estimated to be around 8 nm.
In summary, a template-engaged approach has been demonstrated for generating Se@CdSe nanocables and CdSe
nanotubes as uniform samples and in large quantities. The key
process is the disproportionation of elemental Se into Se2
and SeO32 species at the surfaces of individual Se nanowires,
which led to the formation of a thin (510 nm), uniform, conformal sheath around each template. The CdSe nanoparticles
contained in the sheaths of both nanocable and nanotube
1742

2003 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

structures displayed size-confinement effects, as shown by absorption and photoluminescence measurements. These CdSe
nanostructures with 1D morphology might find use in fabricating optoelectronic devices, and in particular, solar cells
where it has been demonstrated that the efficiency could be
improved by replacing CdSe quantum dots with nanorods.[13]
In addition, the method demonstrated here seems to be a generic one, and could be readily extended to process a range of
binary chalcogenides (e.g., ZnSe, PbSe, and Bi2Se3) into
nanocable and nanotube structures for various applications.[16]

Experimental
In a typical synthesis, freshly prepared selenium nanowires (0.24 g, 3 mmol)
were redispersed in 1 mL ethanol and added to a 25 mL flask containing 15 mL
CdCl2 or Cd(NO3)2 (both purchased from Aldrich) aqueous solution, with the
molar ratio between Cd2+ and Se set at 10:1. The mixture was then heated to
~ 100 C under vigorous magnetic stirring for 8 h. After the reaction mixture
had been cooled down to room temperature, it was centrifuged to yield a
brownblack precipitate. By washing with hot water several times, the by-product, CdSeO3 solid, could be effectively removed from this precipitate to generate pure nanocables in the form of Se@CdSe. Uniform nanotubes of CdSe were
left behind as the final product when the unreacted, t-Se cores were sublimated
by heating the samples at 230 C for 13 min. The diameter of CdSe nanotubes
increased slightly as compared to that of the original templatet-Se nanowires.
If needed, the t-Se nanowires could be briefly sonicated for 3 min to break
them into relatively short nanorods. In this case, cylindrical shells (see Fig. 3D)
of CdSe were obtained as the final product.
TEM images and electron diffraction patterns were taken using a Philips
EM-430 machine operated at 200 kV. The sample was prepared by placing one

http://www.advmat.de

Adv. Mater. 2003, 15, No. 20, October 16

[7]

[8]
[9]

[10]

[11]
[12]
[13]
[14]

[15]
[16]
[17]
[18]

Fig. 4. A) UV-vis extinction spectra and B) photoluminescence spectra, showing


the relative intensity, were taken from solutions containing the following samples: t-Se nanowires, Se@CdSe nanocables, and CdSe nanotubes, respectively.
drop of the suspension on a TEM grid, and letting the solvent evaporate slowly
in a fume hood. XRD patterns were recorded on powder samples using a Philips PW1710 diffractometer (Cu Ka radiation, k = 1.54056 ) at a scanning rate
of 0.02 s1 for 2h in the range of 10 to 70. UV-vis spectra were measured using
a diode array spectrophotometer (Hewlett Packard 8452 A, Palo Alto, CA)
with a resolution of 2 nm. Photoluminescence spectra were recorded using a luminescence spectrophotometer (Perkin Elmer LS-50B, Norwalk, CT) with
pulsed high pressure xenon source.
Received: July 22, 2003
Final version: August 18, 2003

[1]

[2]
[3]
[4]
[5]

Y. Xia, P. Yang, Y. Sun, Y. Wu, B. Mayers, B. Gates, Y. Yin, F. Kim,


H. Yan, Adv. Mater. 2003, 15, 353.
a) S. O. Obare, N. R. Jana, C. J. Murphy, Nano Lett. 2001, 1, 601. b) K. S.
Mayya, D. I. Gittins, A. M. Dibaj, F. Caruso, Nano Lett. 2001, 1, 727.
c) Y. Yin, Y. Lu, Y. Sun, Y. Xia, Nano Lett. 2002, 2, 427.
R. R. He, M. Law, R. Fan, F. Kim, P. Yang, Nano Lett. 2002, 2, 1109.
a) L. J. Lauhon, M. S. Gudiksen, D. Wang, C. M. Lieber, Nature 2002,
420, 57. b) Y. Zhang, K. Suenaga, C. Colliex, S. Iijima, Science 1998, 281,
973.
a) H. Dai, E. W. Wong, Y. Z. Lu, S. Fan, C. M. Lieber, Nature 1995, 375,
769. b) Y. Zhang, H. Dai, Appl. Phys. Lett. 2000, 77, 3015. c) Y. Zhang,
N. W. Franklin, R. J. Chen, H. Dai, Chem. Phys. Lett. 2000, 331, 35.
d) P. M. Ajayan, O. Stephan, P. Redlich, C. Colliex, Nature 1995, 375, 564.
e) C. Tang, S. Fan, M. L. de la Chapelle, H. Dang, P. Li, Adv. Mater. 2000,
12, 1346. f) J. Zhu, S. Fan, J. Mater. Res. 1999, 14, 1175. g) C. N. R. Rao,
B. C. Satishkumar, A. Govindaraj, Chem. Commun. 1997, 1581. h) Y. Wu,
B. Messer, P. Yang, Adv. Mater. 2001, 13, 1487. i) J. H. Song, Y. Wu,

Adv. Mater. 2003, 15, No. 20, October 16

DOI: 10.1002/adma.200305065

PolymerClay Nanocomposite Foams Prepared


Using Carbon Dioxide**
By Changchun Zeng, Xiangmin Han, L. James Lee,*
Kurt W. Koelling, and David L. Tomasko
Polymeric foams (or porous polymeric materials) are used
in many applications because of their excellent strength-toweight ratio, good thermal and sound insulation properties,
flexibility of generating desired morphologies to meet specific
applications, materials savings, etc.[1] Foams with nanometersized voids are under investigation for potential applications
as the next generation materials of low dielectric constants.[2]
However, compared to bulk polymers, foams have reduced
mechanical strength and lower dimensional and thermal stability. Recently developed microcellular foams provide improved mechanical properties over conventional foams,[3] but

[*] Prof. L. J. Lee, C. Zeng, Dr. X. Han, Prof. K. W. Koelling,


Prof. D. L. Tomasko
Department of Chemical Engineering
The Ohio State University, 125 A Koffolt Lab
140 W. 19th Ave., Columbus, OH 43210 (USA)
E-mail: lee.31@osu.edu

[**] The authors gratefully acknowledge financial support from the National
Science Foundation and NSF I/UCR Center for Advanced Polymer and
Composite Engineering (CAPCE) at The Ohio State University.

2003 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

1743

COMMUNICATIONS

[6]

B. Messer, H. Kind, P. Yang, J. Am. Chem. Soc. 2001, 123, 10 397. j) Y.


Sun, B. T. Mayers, Y. Xia, Nano Lett. 2002, 2, 481. k) X. Wen, S. Yang,
Nano Lett. 2002, 2, 451. l) Y. Sun, B. T. Mayers, Adv. Mater. 2003, 15, 641.
m) B. Gates, Y. Wu, Y. Yin, P. Yang, Y. Xia, J. Am. Chem. Soc. 2001, 123,
11 500. n) B. Gates, B. Mayers, Y. Wu, Y. Sun, B. Cattle, P. Yang, Y. Xia,
Adv. Func. Mater. 2002, 12, 679.
a) B. Gates, B. Mayers, A. Grossman, Y. Xia, Adv. Mater. 2002, 14, 1749.
b) B. Mayers, K. Liu, D. Sunderland, Y. Xia, Chem. Mater. 2003, in press.
a) D. M. Chizhikov, V. P. Shchastlivyi, Selenium and Selenides, Collet's,
London and Wellingborough 1968, p. 39. b) C. Wang, W. Zhang, X. Qian,
X. Zhang, Y. Xie, Y. Qian, Mater. Chem. Phys. 1999, 60, 99. c) Q. Peng, Y.
Dong, Z. Deng, Y. Li, Inorg. Chem. 2002, 41, 5249.
D. M. Chizhikov, V. P. Shchastlivyi, Selenium and Selenides, Collet's,
London and Wellingborough 1968, p. 136.
a) C. B. Murray, C. R. Kagan, M. G. Bawendi, Annu. Rev. Mater. Sci.
2000, 30, 545. b) A. P. Alivisatos, J. Phys. Chem. 1996, 100, 13 226. c) H.
Weller, Adv. Mater. 1993, 5, 88. d) J. Aldana, Y. A. Wang, X. Peng, J. Am.
Chem. Soc. 2001, 123, 8844.
a) X. Peng, L. Manna, W. D. Yang, J. Wichham, E. Scher, A. Kadavanich,
A. P. Alivisatos, Nature 2000, 404, 59. b) J. Hu, L. Li, W. Yang, L. Manna,
L. Wang, A. P. Alivisatos, Science 2001, 292, 2060. c) L. Manna, E. C.
Scher, A. P. Alivisatos, J. Am. Chem. Soc. 2000, 122, 12 700. d) Z. Peng,
X. Peng, J. Am. Chem. Soc. 2002, 124, 3343.
M. C. Schlamp, X. Peng, A. P. Alivisatos, J. Appl. Phys. 1997, 82, 5837.
V. I. Kimov, A. A. Mikhailovsky, S. Xu, A. Malko, J. A. Hollingsworth,
C. A. Leatherdale, H. J. Eisler, M. G. Bawendi, Science 2000, 290, 314.
W. U. Huynh, J. J. Dittmer, A. P. Alivisatos, Science 2002, 295, 2425.
a) W. C. W. Chan, S. M. Nie, Science 1998, 281, 2016. b) M. Bruchez,
M. Moronne, P. Gin, S. Weiss, A. P. Alivisatos, Science 1998, 281, 2013.
c) B. Dubertret, P. Skourides, D. J. Norris, V. Noireaux, A. H. Brivanlou,
A. Libchaber, Science 2002, 298, 1759.
C. N. R. Rao, A. Govindaraj. F. L. Deepak, N. A. Gunari, Appl. Phys.
Lett. 2001, 78, 1853.
a) B. L. Wehrenberg, C. Wang, P. Guyot-Sionnest, J. Phys. Chem. B 2002,
106, 10 634. b) A. Al Bayaz, A. Giani, M. C. Artaud, A. Foucaran, F. Pascal-Delannoy, A. Boyer, J. Cryst. Growth 2002, 241, 463.
B. T. Kolomiets, A. A. Malkov, Fiz. Tverd. Tela 1959, 1, 32.
C. B. Murray, D. J. Norris, M. G. Bawendi, J. Am. Chem. Soc. 1993, 115,
8706.

You might also like