Faculty & Staff Scholarship
2005
Size-Dependent Properties Of Cdse Quantum Dots Size-Dependent Properties Of Cdse Quantum Dots
S. Neeleshwar
C. L. Chen
C. B. Tsai
Y. Y. Chen
C. C. Chen
See next page for additional authors
Follow this and additional works at: https://researchrepository.wvu.edu/faculty_publications
Digital Commons Citation Digital Commons Citation
Neeleshwar, S.; Chen, C. L.; Tsai, C. B.; Chen, Y. Y.; Chen, C. C.; Shyu, S. G.; and Seehra, M. S., "Size-
Dependent Properties Of Cdse Quantum Dots" (2005).
Faculty & Staff Scholarship
. 676.
https://researchrepository.wvu.edu/faculty_publications/676
This Article is brought to you for free and open access by The Research Repository @ WVU. It has been accepted
for inclusion in Faculty & Staff Scholarship by an authorized administrator of The Research Repository @ WVU. For
more information, please contact [email protected].
Authors Authors
S. Neeleshwar, C. L. Chen, C. B. Tsai, Y. Y. Chen, C. C. Chen, S. G. Shyu, and M. S. Seehra
This article is available at The Research Repository @ WVU: https://researchrepository.wvu.edu/faculty_publications/
676
Size-dependent properties of CdSe quantum dots
S. Neeleshwar, C. L. Chen, C. B. Tsai, and Y. Y. Chen
*
Institute of Physics, Academia Sinica, Taipei, Taiwan, Republic of China
C. C. Chen
Department of Chemistry, National Taiwan Normal University, and Institute of Atomic and Molecular Sciences,
Academia Sinica, Taipei, Taiwan, Republic of China
S. G. Shyu
Institute of Chemistry, Academia Sinica, Taipei, Taiwan, Republic of China
M. S. Seehra
Physics Department, West Virginia University, Morgantown, West Virginia 26506, USA
Received 16 March 2005; revised manuscript received 27 April 2005; published 23 May 2005
Temperature dependences of the magnetic susceptibility
and heat capacity C
p
of CdSe quantum dots with
size d=2.8, 4.1, and 5.6 nm are compared to those of bulk CdSe to determine the size-dependent effects. With
decreasing size d, the following effects are observed: iroom temperature optical absorption shows a blueshift
of the band gap; ii room temperature x-ray diffraction show wurtzite structure but with smaller lattice
constants; iii magnetic susceptibility changes from negative diamagnetic for the bulk to positive
with
magnitude increasing with decreasing d; and iv the Sommerfeld constant
determined from the C
p
/T vs T
2
data increases. Possible explanations for these size-dependent properties are presented.
DOI: 10.1103/PhysRevB.71.201307 PACS numbers: 75.75.a, 65.80.n
I. INTRODUCTION
In recent years, properties of nanosized materials have
generated a great deal of interest because of the science in-
volved in these studies and technological applications of the
quantum dots QDs. As the physical dimensions of the par-
ticle approach to the nanometer scales, quantization and sur-
face effects begin to play an important role, leading to drastic
changes in measured properties.
1
Among the semiconductor
QD, studies have been reported for the II-IV Ref. 2 and
III-V Ref. 3 materials, where a shift in the electronic tran-
sitions to higher energies accompanied by an increase of the
oscillator strength with the decrease in the particles size were
reported. Applications of the semiconductor QD have been
reported for photovoltaics,
4
light emitting diodes,
5
lasers,
6
and biological imagings.
7
Other reports studied include opti-
cal spectroscopy,
8
photoconductivity,
9
and LO-phonon
coupling.
10
None of the studies listed above in semiconductor QD
have focused on the effect of size on thermodynamic prop-
erties such as magnetic susceptibility
and heat capacity C
p
.
Consequently in this paper we report detailed studies of the
temperature dependence of
and C
p
for CdSe quantum dots
with size d=2.8, 4.1, and 5.6 nm vis-a-vis bulk CdSe. Im-
portant size-dependent effects are observed, whose discus-
sion and analysis are presented below.
II. EXPERIMENTAL
CdSe semiconductor quantum dots were prepared from
the pyrolysis of dimethylcadmium and tri-n-octylphosphine
selenide TOPSe in a hot coordinating solvent of tri-
n-octylphosphine oxide TOPO using the procedure de-
scribed previously.
11
In this method, the surface of the CdSe
quantum dot was passivated with TOPO molecules to avoid
surface oxidation and aggregation. Different sizes of quan-
tum dots were obtained by controlling its nucleation and
growth process. For further size selection, size-selective pre-
cipitation can be carried out in a chloroform-methanol sol-
vent system. Three sizes of quantum dots were prepared with
d=2.8, 4.1, and 5.6 nm with a standard deviation of 10%
as determined by the high-resolution transmission electron
microscopy HRTEM; see inset of Fig. 1 for the 5.6-nm QD.
Optical absorption spectra of CdSe quantum dots were ob-
tained by a HP 8452 diode array spectrophotometer using
FIG. 1. Optical absorption spectra for d=2.8-, 4.1-, and 5.6-nm
CdSe quantum dots dispersed in chloroform were taken at 300 K.
Inset: The HRTEM image of 5.6-nm CdSe quantum dots; an ex-
ample particle marked by a circle is shown.
PHYSICAL REVIEW B 71, 201307R兲共2005
RAPID COMMUNICATIONS
1098-0121/2005/7120/2013074/$23.00 ©2005 The American Physical Society201307-1
1-cm quartz cuvettes at room temperature as shown in Fig. 1.
The blueshift of the absorption edge with the decreasing d of
quantum dots is consistent with an earlier report.
12
X-ray
diffraction XRD of the quantum dots carried out with a
3-KW Philips diffractometer equipped with an array detector
based on a real-time multiple strip showed the wurtzite
structure of the bulk CdSe but with the expected line broad-
ening with decreasing d Fig. 2. In addition, there is a
shrinkage of the lattice constants the inset of Fig. 2, due to
size effect, somewhat similar to that reported in the
literature.
13
No additional lines due to any impurity phase
could be detected in the XRD spectra.
A calorimetric study was made in the range of 0.4 to 10
K, using a thermal-relaxation microcalorimeter in a
3
He
cryostat.
14
Each milligram-sized sample was prepared by
lightly pressing fine powders together. It was then attached
with thermal-conducting N grease to a sapphire disk, having
two deposited thin films serving as heater and thermometer,
respectively. The heat capacity of the sapphire disk and
grease were measured separately, and used as addenda cor-
rection in data analysis. The relative precision and the abso-
lute accuracy of the calorimeter were confirmed to be within
3% by measuring the copper standard. Magnetization mea-
surements were performed as a function of temperature using
the Quantum Design superconducting quantum interference
device SQUID magnetometer in the range 2 to 300 K. The
magnetic susceptibility of straw and capsule were measured
separately and subtracted from the data.
III. RESULTS AND DISCUSSION
The temperature dependence of magnetic susceptibility
for CdSe QD with size d=2.8, 4.1, and 5.6 nm and bulk
CdSe is shown in Fig. 3. For the bulk CdSe,
is diamagnetic
and temperature independent with the magnitude
−4310
−6
emu/mole in good agreement with the earlier
results.
15
We report that for QD,
is positive and it has a
strong temperature dependence, especially below 30 K. Also
the magnitude of
is larger for the smaller particles at all
temperatures, showing the effect of size on magnetism.
In general for pure semiconductors,
=
l
+
f
+
i
, where
l
is the temperature-independent lattice contribution,
f
is
the free charge carrier electrons and holes contribution and
i
is the contribution from bounded carriers and dangling
bonds. For bulk CdSe,
f
,
i
l
, leading to magnetic sus-
ceptibility determined by
l
, which is usually negative
Fig. 3.
15
Shaldin et al.
16
have shown that in II-IV semicon-
ductors, vacancies and interstitial can occur during the
growth. For QD, such defects will be more prevalent as com-
pared to bulk materials because of the increase in the relative
surface area. Specific magnetic clusters created by the donor-
acceptor pairs can exhibit paramagnetic behavior.
17
On the
surfaces of semiconductors, the free dangling bond bears an
electron spin by nature and can make semiconductor surfaces
magnetic. These phenomena are expected to be more signifi-
cant in QD.
18
With these considerations in mind, we suggest that the
low-temperature Curie tail in
is most likely due to surface
dangling bonds. These surface dangling bonds result from
decreased coordination of the surface atoms of the QD. We
have fitted the low-temperature data for T 30 K to the
modified Curie law:
=
o
+C/T, where
o
is temperature
independent contribution mainly from Pauli paramagnetism
of
f
mentioned above. The details will be discussed later.
The fits are excellent with C=1.33, 4.0, and 8.38 in units of
10
−4
emu K/mol for d=5.6, 4.1, and 2.8 nm, respectively
inset to Fig. 3. This rapid increase in C with a decrease in
d is due to increase in surface/volume ratio as d decreases.
Note that C=N
2
/3k
B
where N is the number of dangling
bonds/mol, each with effective magnetic moment
and
k
B
is the Boltzmann constant. If we assume spin S=1/2
with each dangling bond, leading to
=1.73
B
, then
N=13.5 10
20
/mol for d=2.8 nm, thus yielding the concen-
tration of the dangling bonds 2000 ppm. For d=4.1 and
5.6 nm, a similar calculation yields the concentration 1000
and 300 ppm, respectively. It is noted that in amorphous Si
and Ge, low-temperature magnetic susceptibility studies
yielded similar concentration of spin density due to dangling
bonds.
19
The increase in
with increasing temperature above 30 K
seen for the QD in Fig. 3 is another interesting feature of our
FIG. 2. X-ray diffraction patterns for the bulk and d=2.8-, 4.1-,
and 5.6-nm quantum dots. Inset: The size dependence of lattice
constants of a and c axes.
FIG. 3. The magnetic susceptibility as function of temperature
for the bulk and d=2.8-, 4.1-, and 5.6-nm quantum dots; the lines
are for eye’s guide. Inset: The Curie constant vs d.
NEELESHWAR et al. PHYSICAL REVIEW B 71, 201307R兲共2005
RAPID COMMUNICATIONS
201307-2
results. At the outset we note that a similar increase was
reported by Burgardt and Seehra in semiconductor FeS
2
.
20
In
Fig. 3, both the magnitude and the slope increase with de-
crease in d. For FeS
2
Ref. 20 and amorphous Si and Ge,
21
the positive
and its temperature dependence at higher tem-
perature were explained by the Van Vleck susceptibility
vv
=2N
A
B
2
k
兩具lL
Z
k典兩
2
E
k
E
l
, 1
where N
A
is the Avogadro’s number and L
Z
is the z compo-
nent of the orbital angular momentum coupling the excited
state k with energy E
k
with the ground state l with energy
E
l
. For semiconductors, E
k
E
l
E
g
energy gap. Note that
E
g
usually decreases with increase in temperature
2,22
and in
CdS nanoclusters, a much steeper temperature dependence
with decreasing particle size is observed. Assuming similar
results are valid for CdSe QD, it then explains why
in-
creases with increasing temperature, and increasing slope
with decreasing d, as observed in Fig. 3. To estimate
vv
from Eq. 1, if we approximate the sum over all the states
by 1/E
g
assuming the matrix elements to be unity,
vv
=0.3 10
−4
emu/mol Oe is obtained for E
g
1.75 eV
valid for CdSe QD. This estimate of
vv
is about a factor of
three times smaller than the enhancement of
observed for
QD. This confirms that
o
is mainly contributed by
f
as
proposed earlier. This issue requires further investigation.
Since the surface free charge carriers which gives
f
are
easily formed in QDs,
13
the increase in the number of free
charge carriers with surface for smaller particles is expected
to vary as 1/d. In Fig. 4
o
vs 1/d shows linear dependence.
The fact that there are systematic changes, in both the mag-
nitudes and temperature dependence of
, with the particle
size d suggests that
is dominated by the size effect and
surface effects rather than any impurity.
To further examine the consequences of the size effect,
measurements of specific heat of bulk CdSe and quantum
dots with size d=2.8, 4.1, and 5.6 nm were carried out for
T=0.340 K. The temperature dependence of specific heat
for the bulk CdSe and quantum dots, plotted as C/T vs T
2
is
shown in Fig. 5. The heat capacity of the bulk is in good
agreement with earlier report.
23
The enhancement of specific
heat of quantum dots as the evolution of size is clearly re-
vealed. In general, the specific heat of a material can be
represented by the summation of contributions of conduction
electrons C
el
=
T, lattice phonon C
ph
and magnetic correla-
tions C
mag
. The value of the Sommerfeld constant
obtained
from the intercept of the linear fits gradually increases from
1.12 mJ/mole K
2
for the bulk to 5.50 mJ/mole K
2
for 2.8
nm with decreasing d. The relation of
and the density of
states of conduction electrons N
F
can be represented by
=
1
3
2
k
B
2
N
F
, 2
where k
B
is Boltzmann constant. The value of
increases
with decreasing d and indicates an enhancement of density of
states of conduction electrons N
F
in quantum dots. The
value of
is approximately linear proportional to 1/d, im-
plying the correlation of the density of states of conduction
electrons N
F
with the surface of quantum dot Fig. 4.
Since
f
is also proportional to N
F
, the similar variations
of
and
o
are understandable. It is noted that quantum dots
have an enormous surface-to-volume ratio; consequently, the
free charge from delocalized electrons of dangling bonds and
defects on surface will have more contribution to magnetic
susceptibility
o
and heat capacity
as well. For quantum
dots, the lattice phonons C
ph
can be calculated by the theo-
retical model for a small particle represented by the follow-
ing equation:
24
C
ph
= V
m
l,s
32l +1k
B
x
2
e
x
4
R
3
e
x
−1
2
, with x =
ca
l,s
RT
. 3
Here V
m
is the molar volume, R denotes the particle ra-
dius, a
l,s
,isthesth zero of the derivative of the lth spherical
Bessel function, and c is the effective sound velocity. The
number of atoms N
o
in quantum dots with d=2.8, 4.1, and
5.6 nm is estimated to be about 500, 1200, and 2200, respec-
tively. We use the constraint
l
l max
2l+1=N
o
and subtract
the contribution to the heat capacity from free charge carriers
C
el
=
T. The remaining heat capacity yields C
ph
from which
c=795, 895, 915 m/s with Debye temperature =61, 68, 70
K are obtained for d=2.8, 4.1, and 5.6 nm, respectively.
FIG. 4. The Sommerfield constant and
o
mainly from the con-
tribution of free charge carrier vs 1/d 共⬃surface/volume ratio,
the lines are linear fits.
FIG. 5. The specific heat, plotted as C/T vs T
2
for the bulk and
d=2.8-, 4.1-, and 5.6-nm quantum dots; the lines are linear fits.
SIZE-DEPENDENT PROPERTIES OF CdSe QUANTUM DOTS PHYSICAL REVIEW B 71, 201307R兲共2005
RAPID COMMUNICATIONS
201307-3
Compared to =139 K for bulk CdSe, for the quantum
dots are really half, an anticipated result from lattice soften-
ing with decreasing d.
24
IV. CONCLUSION
The optical absorption spectra show a blueshift in the
CdSe quantum dot. X-ray diffraction confirmed that QDs
have the same wurtzite crystal structure as the bulk but with
smaller lattice constants. The low-temperature magnetic sus-
ceptibility studies reveal the increase of spin density of dan-
gling bonds with decreasing size. The magnetic susceptibility
o
and Sommerfeld constant
increases linearly with
surface-to-volume ratio, giving the evidence of free charge
carriers on the surface of CdSe quantum dot. The systematic
changes in the magnitudes of
and
with the size d suggest
the role of quantum size effect and surface effects rather than
any impurity in CdSe quantum dots.
ACKNOWLEDGMENT
This work was supported by the National Research Coun-
cil of the Republic of China under Grant No. NSC 93-2112-
M-001-022.
*
Electronic address: [email protected]
1
A. P. Alivisatos, Science 271, 933 1996.
2
T. Vossmeyer, L. Katsikas, M. Giersig, I. G. Popovic, K. Diesner,
A. Chemseddine, A. Eychmüller, and H. Weller, J. Phys. Chem.
98, 7665 1994.
3
A. A. Guzelian, U. Banin, A. V. Kadavanich, X. Peng, and A. P.
Alivisatos, Appl. Phys. Lett. 69, 1432 1996.
4
R. P. Raffaelle, S. L. Castro, A. F. Hepp, and S. G. Bailey, Prog.
Photovoltaics 10, 433 2002.
5
S. Coe, W. K. Woo, M. G. Bawendi, and V. Bulovic, Nature
London 420, 800 2002.
6
V. I. Klimov, A. A. Mikhailovsky, S. Xu, A. Malko, J. A. Holl-
ingsworth, C. A. Leatherdale, H. J. Eisler, and M. G. Bawendi,
Science 290, 314 2000.
7
M. Bruchez, Jr., M. Moronne, P. Gin, S. Weiss, and A. P. Alivi-
satos, Science 281, 2013 1998.
8
M. Nirmal, C. B. Murray, and M. G. Bawendi, Phys. Rev. B 50,
2293 1994.
9
M. C. Beard, G. M. Turner, and C. A. Schmuttenmaer, Nano Lett.
2, 983 2002.
10
M. Nirmal, C. B. Murray, D. J. Norris, and M. G. Bawendi, Z.
Phys. D: At., Mol. Clusters 26, 361 1993.
11
C. B. Murray, D. J. Norris, and M. G. Bawendi, J. Am. Chem.
Soc. 115, 8706 1993.
12
C. D. Dushkin, S. Saita, K. Yoshie, and Y. Yamaguchi, Adv. Col-
loid Interface Sci. 88,372000.
13
C. B. Murray, C. R. Kagan, and M. G. Bawendi, Annu. Rev.
Mater. Sci. 30, 545 2000.
14
Y. Y. Chen, Y. D. Yao, S. S. Hsiao, S. U. Jen, B. T. Lin, H. M.
Lin, and C. Y. Tung, Phys. Rev. B 52, 9364 1995.
15
D. J. Chadi, R. M. White, and W. A. Harrison, Phys. Rev. Lett.
35, 1372 1975.
16
Y. V. Shaldin, I. Warchulska, M. K. Rabadanov, and V. K. Komar,
Semiconductors 38, 288 2004.
17
J. van Wieringen, Philips Tech. Rev. 19, 301 1957/1958.
18
Takanori Suzuki, V. Venkataramanan, and Masakazu Aono,
RIKEN Rev. 37,92001.
19
S. J. Hudgens, Phys. Rev. B 14, 1547 1976.
20
P. Burgardt and M. S. Seehra, Solid State Commun. 22, 153
1977.
21
S. J. Hudgens, Phys. Rev. B 7, 2481 1973.
22
M. S. Seehra and S. S. Seehra, Phys. Rev. B 19, 6620 1979.
23
A. Twardowski, H. J. M. Swagten, and W. J. M. de Jonge, Phys.
Rev. B 42, 2455 1990.
24
G. H. Comsa, D. Heitkamp, and H. S. Räde, Solid State Commun.
24, 547 1977.
NEELESHWAR et al. PHYSICAL REVIEW B 71, 201307R兲共2005
RAPID COMMUNICATIONS
201307-4