You are on page 1of 7

This article appeared in a journal published by Elsevier.

The attached
copy is furnished to the author for internal non-commercial research
and education use, including for instruction at the authors institution
and sharing with colleagues.
Other uses, including reproduction and distribution, or selling or
licensing copies, or posting to personal, institutional or third party
websites are prohibited.
In most cases authors are permitted to post their version of the
article (e.g. in Word or Tex form) to their personal website or
institutional repository. Authors requiring further information
regarding Elseviers archiving and manuscript policies are
encouraged to visit:
http://www.elsevier.com/copyright

Author's personal copy

Available online at www.sciencedirect.com

Optical Materials 30 (2008) 14401445


www.elsevier.com/locate/optmat

Linear and nonlinear optical refractions of CR39 composite


with CdSe nanocrystals
Chenli Gan a, Yanpeng Zhang a, S.W. Liu a, Yunjun Wang b, Min Xiao
a

a,*

Department of Physics, University of Arkansas, Fayetteville, AR 72701, United States


b
Mesolight LLC, Fayetteville, AR 72701, United States

Received 21 May 2007; received in revised form 6 July 2007; accepted 30 August 2007
Available online 23 October 2007

Abstract
Hybrid composites of CdSe nanocrystals embedded in allyl diglycol carbonate (CR39) matrices have been prepared and characterized.
The measurements show that the linear refractive index of the composite decreases as the CdSe nanocrystals weight-percentage concentration increases at the laser wavelengths of 632.8 nm and 532 nm. The room temperature nonlinear optical properties of the hybrid composites were investigated using a single-beam Z-scan technique with femtosecond laser pulses at the wavelengths of 794 nm and 397 nm.
The experimental data reveals that the Kerr nonlinear refractive index n2 of the composite increase at these wavelengths when the CdSe
nanocrystals weight-percentage concentration increases. Also, the nonlinear refractive index n2 of the CdSe/CR39 hybrid composites
exhibit dispersion from a positive value at 794 nm (below the band gap) to a positive value at 395 nm (above the band gap). The measured dispersion of n2 is roughly consistent with the Sheik-Bahaes theory for the bound electronic nonlinear refraction resulting from the
two-photon resonance.
Published by Elsevier B.V.
PACS: 42.65
Keywords: CdSe/CR39 lter; Nonlinear refraction index

1. Introduction
Owing to their unique size-dependent optical and electrical properties, colloidal semiconductor nanocrystals are
considered as a class of very promising materials for the
next generation of electronic and optoelectronic devices
such as light emitting diodes (LEDs), solar cells, lasers,
printing of integrated circuits, biological detection, etc. A
major advantage of using nanocrystals is that they can be
embedded into organic, polymeric, and even glass matrices
to make nanocomposite materials. In particular, the nanocrystal/polymer composites have the combined advantages
of inorganic and organic materials, e.g. low-production

Corresponding author. Tel.: +1 479 5756568; fax: +1 479 5754580.


E-mail address: mxiao@uark.edu (M. Xiao).

0925-3467/$ - see front matter Published by Elsevier B.V.


doi:10.1016/j.optmat.2007.08.010

cost and exible possibilities, etc. However, studies on the


refractive properties of those nanocrystal/polymer nanocomposite materials are very limited.
Traditional optical lters rely on light absorption, interference or both [1]. Semiconductors that exhibit a very
rapid transition from opacity to transparency edge are particularly useful for such application, making excellent longpass optical lters. However, the xed edges of absorption
(band gap energy) of a given semiconductors signicantly
limit their applications. On the other hand, the designs
and fabrications of interference-based bandpass lters
can be a very complicated and expensive process. Therefore, there is a need to develop new types of nanocrystalbased polymer thin-lm optical lters that can be easily
fabricated with high quality stability, and tunable wavelength. The composite lters with CdSe nanocrystals
embedded in polymer (CR39) matrix described in this work

Author's personal copy

C. Gan et al. / Optical Materials 30 (2008) 14401445

have a weak nonlinearity. The third-order susceptibility of


these composite materials is two-orders of magnitude smaller than the CdSe-doped inorganic glass lters [2]. Also,
such system can be easily expanded to other semiconductor
nanocrystals like CdS and CdTe, since they have similar a
nonlinear refractive index [27a].
Taking advantage of the unique size-dependent feature
of semiconductor nanostructures, we developed a technique of homogenously dispersing CdSe nanocrystals into
CR39 polymers and then fabricated thin lms using an
in situ polymerization process instead of the well-established polymer processing technique such as extrusion or
heat and press process. We performed optical characterizations of such nanocrystal-based thin lm lters, including measuring linear refraction index and nonlinear
refraction index by single-beam Z-scan technique. Such
detailed studies are important in developing practical
nanocrystal-based optical lters with improved optical
characteristics. For high power applications of such nanocrystal/polymer optical lters, the nonlinear optical characteristics need to be investigated. Also, such nonlinearity
might be used to make optical limiters for applications in
certain wavelength regions.
2. Sample preparation
CR39 is a plastic polymer that is widely used in manufacturing eyeglass lenses. It is transparent in visible spectrum and almost completely opaque in infrared and
ultraviolet wavelength ranges. Because of the presence of
the two allyl functional groups (Fig. 1), CR39s monomers
can not only polymerize but also cross-link with each other
which result in a thermoset plastic characterized by being
hard, infusible and insoluble in all solvents. In addition,
CR39 homopolymer is a high-grade optical plastic whose
linear index of refraction is just slightly less than that of
the crown glass.
Diisopropyl peroxydicarbonate (IPP) is usually used to
catalyze CR39s polymerization. However, due to its extremely pyrophoric nature, commercial IPP is only available
for industrial uses but not for research purpose. Because
of this, an alternative catalyst, benzyl peroxide, is used to
initialize the polymerization of CR39.
Since CR39 is resistant to most of the organic solvents,
an in situ polymerization strategy is used to prepare the
composite material. To do this, various ligands were
screened to identify the one that would make the nanocrystals be compatible with the CR39 monomers, and a proper

Fig. 1. CR39s momoner:diethyleneglycol bis allycarbonate.

1441

one was synthesized to be compatible with the CR39


monomer. This ligand contains a thiol functional group
(to bind on nanocrystal surface) and a long polyethyleneglocol (PEG) tail, which, we believe, makes the semiconductor CdSe nanocrystals compatible with the CR39
polymer. The CdSe nanocrystals with the particle size of
3.5 nm were synthesized according to the published
method [7b], and then were thoroughly puried by precipating and washing with methanol and stored in toluene.
An in situ polymerization process was used to form CdSe
nanocrystal/CR39 polymer nanocomposite thin lms. To
form CR39 monomer-compatible CdSe nanocrystals,
native trioctylphosphine and trioctylphosphine oxide
(TOP/TOPO) capped CdSe nanocrystals were mixed with
an excess of poly (ethylene glycols)-terminated dihydrolipoic acid (DHLAPEG) ligand. After being incubated
for several hours at room temperature, the DHLAPEG
coated CdSe nanocrystals were separated from solution
via centrifuging and then re-dispersed into CR39 liquid
monomer to form a homogenous solution (via overnight
stirring). The CdSe nanocrystal/CR39 monomer mixture
were injected into lm molders and cured at temperature
up to 100 C with benzyl peroxides as initiators by a classical curing process of free-radical-addition polymerization
[7c]. Fig. 2 (color online) shows the CdSe/CR39 thin lms
with various nanocrystal concentrations.

3. Optical measurements
3.1. Linear absorption coecient
The UVvis absorption spectra of all six samples with
dierent concentrations of CdSe nanocrystals are presented
in Fig. 3. The pure CR39 sample (sample a) was found to
absorb only below 400 nm, hence the absorption between
400 and 700 nm in samples bf can be attributed to the
added CdSe nanocrystals. It is known that the UV absorption of semiconductor nanocrystals is due to the band gap
absorption, which is blue-shifted relative to the bulk due to
quantum size connement eect [8,9]. Thus, Fig. 3 shows
that the UV thresholds of all samples are blue-shifted relative to bulk CdSe at wavelength 660 nm. It is also noted
that, within the same batch of samples, the absorption
threshold becomes more blue-shifted as the CdSe concentration is decreased. The average particle size in our samples (calculated from the equation based on the eective
mass approximation [8], and compared to the experimental
curve of absorption threshold versus particle diameter by

Fig. 2. (Color online) CdSe nanocrystal/CR39 lms (1.0 mm), the CdSe
concentrations of each lm are labeled on the top of the gure.

Author's personal copy

1442

C. Gan et al. / Optical Materials 30 (2008) 14401445

tions in polymers CR39 to be 0%, 0.16%, 0.34%, 0.5%,


0.92% and 1.48%, respectively.

2.5

Absorbtance (arb)

2.0

3.2. Linear refractive index


1.5

f
e

1.0

d
c
b

0.5

a
0.0

300

400

500

600

700

wavelength(nm)
Fig. 3. Absorption spectra of CR39 lms (thickness = 1.0 mm) embedded
with (a) 0%, (b) 0.16%, (c) 0.34%, (d) 0.5%, (e) 0.92%, and (f) 1.5% (w/w)
of CdSe nanocrystals.

Henglein et al. [9]) is estimated to be about 3.5 nm that is


consistent with the diameter measured by TEM image.
We found that phase segregation occurs at high nanocrystal concentration, e.g., above 1.0%.
The linear absorption coecient a of the CR39 thin
lms embedded with dierent concentrations of CdSe
nanocrystals are shown in Fig. 4 from the transmittance
spectra using the relation
a ln1=T a =l;

where Ta is the interference-free transmittance [10] of the


thin lm after subtracting the substrates absorption, and
l is the thickness of thin lm. The linear absorption coecient a of the composite lm depends on CdSe nanocrystal
concentration in polymer CR39 at a given wavelength,
higher concentration makes a to be larger. The lms have
very small a value in the infrared wavelength region. The
linear absorption coecients (a) of the composite lms drop
sharply in the visible region and the zero edges of a are located at dierent wavelengths 290 nm, 340 nm, 420 nm,
480 nm, 570 nm, 585 nm for CdSe nanocrystal concentra-

By using a prism coupler system, we measured the linear


optical refractive indices of these samples, which also
depend on the concentration of CdSe nanocrystals in
CR39 polymer. It decreases from 1.5066 to 1.4876 with
increasing concentration of CdSe nanocrystals at CR39
polymer when the laser wavelength is 632.8 nm, and it
decreases from 1.5122 to 1.4944 with increasing concentration of CdSe nanocrystals when the laser wavelength is
532 nm. As an example, Fig. 5 shows the variation of the
refractive index versus CdSe nanocrystal concentration
for these CdSe nanocrystal/CR39 polymer lms. As shown
in this gure, the refractive index decreases as the CdSe
nanocrystal weight-percentage concentration increases.
The linear refractive index of CdSe nanocrystals of
3.5 nm average diameter is calculated to be 2.34 with
the following relation [11]
nCdSe f1 ebulk  1=1 0:75 nm=d1:2 g1=2 ;

where ebulk  6.2 is the dielectric constant of bulk CdSe and


d is the average diameter of CdSe nanocrystals in
nanometer.
Thus, the refractive index of the dopant semiconductor
nanocrystals (2.34) is higher than that of the host polymer
(1.5) in our samples, one would naturally anticipate that
embedding nanocrystals into polymer matrix would cause
the increase in the refractive index [12,13]. The observed
slight decrease of n0 with increasing nanocrystal concentration (shown in Fig. 5) may be attributed to the poor crystalline structure due to relaxation rate dierence between
nanocrystals and the polymer host [13]. The decreased
trend in n0 as increasing nanocrystal concentration could
also be due to the change in polymers morphology upon
embedding nanocrystals. This phenomenon is currently
still under investigation.

1.515

Refractive Index (n)

Absorbtion Constant (cm-1)

1.510
532 nm

1.505

633 nm

1.500
1.495
1.490

1.485
0
0
0.0

0.5

1.5

CdSe Weight Percentage


0.4

0.8

1.2

1.6

CdSe Weight Percentage


Fig. 4. Absorption constant of CR39 thin lms (thickness = 1.0 mm)
embedded via dierent CdSe nanocrystals at wavelength 400 nm.

Fig. 5. Variation of the refractive index of CdSe/CR39 nanocomposite


lms against CdSe concentration (weight-percentage). CdSe nanocrystals
used have absorption edge of 570 nm; measurements are performed at two
wavelengths, 532 nm and 633 nm.

Author's personal copy

C. Gan et al. / Optical Materials 30 (2008) 14401445


1.0

The quantum connement and dielectric connement


eects make semiconductor nanocrystals a promising class
of nonlinear optical materials [1417]. We performed the
nonlinear optical measurements of the CdSe/CR39 composite samples by using the single-beam Z-scan technique
[18]. The nonlinear refractive index at both the wavelengths
of 794 nm and 397 nm were measured to show dispersion
properties. The light source is a mode-locked Ti:Sapphire
laser with a high repetition rate of 81 MHz at the wavelength of 794 nm, which has a temporal width of about
565 femtosecond and a peak power of up to 40 kW. The
397 nm pulse laser beam was generated by frequency doubling the 794 nm laser pulses with a peak power of up to
4.9 kW. The output beam was collimated, expanded and
the spatial proles of these laser beams were determined
to be approximately Gaussian. The beam was then focused
by a lens with the focal waist radius of x0  11 lm
(z0 = 0.5 mm) at 794 nm (or 7 lm (z0 = 0.4 mm) at
397 nm). To reduce the eect of uctuations in the pulse
energy, another detector was used to monitor the input
pulse energy as a reference. All closed-aperture (CA) Zscan measurements were performed with a xed linear
transmittance of S = 4%, dened as the fraction of the
beam energy that passed through the far-eld aperture.
The open-aperture (OA) and CA Z-scan traces are tted
by two corresponding numerical programs based on Ref.
[19] to obtain the nonlinear refractive indices. The samples
parameters (a linear optical absorption coecient and L
sample thickness), as well as the lasers parameters (R repetition rate, tFWHM FWHM of pulses duration, and AP
average incident power of the laser beam) are required as
the inputs for these numerical programs.
We can use the equations [18] below to calculate the
nonlinear refraction indexes n2 of the lms by the Z-scan
traces as following

0.9

n2

DT pv

0:4061  S0:25 kLeff I 0


2p
;
k
k
Leff 1  eaL =a;

;
3

where L is the sample thickness, a is the linear absorption


coecient, and DTpv is the peak-valley transmittance
change.
A typical CA Z-scan curve of the CdSe/CR39 thin lm
(with CdSe weight concentration of 0.16%) is shown in
Fig. 6, which gives the characteristic shape for a positive
nonlinearity. Assuming that the nonlinear refractive index
of this sample is dominated by the third-order only, the
experimentally measured data (solid line) is tted with
the standard Z-scan theory to extract the nonlinear refractive index. The measurement in Fig. 6 was done at a xed
input irradiance with the laser wavelength of 397 nm and
the pulse duration of 565 fs. From the tting results the

Normalized Transmitance

3.3. Nonlinear refractive index

1443

0.8
0.7
0.6
0.5
0.4
-1.5

-1.0

-0.5

0.0

0.5

1.0

1.5

Z Position (cm)
Fig. 6. Measured Z-scan CA trace in CdSe/CR39 lm (1.0 mm) at
wavelength 397 nm. CdSe weight-percentage: 0.16%; S = 0.04; n2 = 1.09
104 cm2/GW.

observed nonlinear refractive index is determined to be


about 1.0 104 cm2/GW, which is of the same magnitude
as the ones reported in Refs. [2,3]. Seo et al. [3] measured
the third-order nonlinear optical susceptibility of CdSe
nanocrystals to be 1.8 1051.8 104 cm2/GW for various concentrations of CdSe quantum dots (1.2 106
9.9 103 mol/m3) in a toluene host at the wavelength of
532 nm by using nanosecond degenerate four-wave mixing
with 8 ns laser pulses. Yao et al. [2] got the value of
1.1 107 esu (9.1 104 cm2/GW) for jv(3)j of CdSedoped PHEMA lms (the absorption coecient of the lm
is 20.7 cm1 at 460 nm), indicating low concentration of
doped nanocrystals. The reectivity of their signal was
about two-orders of magnitude smaller than that for
CdSe-doped inorganic glasses [4].
Also the nonlinear refractive index of the CdSPS
hybrid composites was found to increase linearly (in the
range from 1.0 104 cm2/GW to 3.0 104 cm2/
GW) as the input irradiance decreases. These values are
of the same magnitude as the results obtained in Ref. [6],
in which the nonlinear refractive index of capped CdS
nanocrystals in an ordered polydiacetylene host at wavelength of 532 nm was measured to be about
1.1 104 cm2/GW using 0.09 GW/cm2 laser pulses of
5 ns. In that case, the polydiacetylene host itself has a nonlinear refractive index of about 3 104 cm2/GW. In
addition, Wang et al. [7] had performed the pump-probe
measurements with ammonia-passivated CdS clusters in
polymer. They measured the absorption of CdS at dierent
wavelengths with excitation by 30 ps laser pulses at 355 nm
and obtained the associated changes in refractive index at
dierent wavelengths through the KramersKronig relation. According to their calculation, the nonlinear refractive index of CdS at 532 nm is 1.41 104 cm2/GW,
which is also in the same range as our results presented
here. In our composite samples, we can attribute the nonlinearity to the CdSe nanoparticles, since the pure CR39
sample dose not show third-order nonlinear properties in
the error range of our measurements.

Author's personal copy

C. Gan et al. / Optical Materials 30 (2008) 14401445

Sheik-Bahae et al. [19] and Krauss and Wise [20] have


presented a scaling rule between the nonlinear refraction
index n2 of hybridized CdSe/CR39 nanocomposites and
the ratio of the photon energy to the band gap energy
(hm/Eg). Here we found that if the mechanism for the nonlinearity in our samples is dominated by the third-order
nonlinear eect, we should be able to t the n2 value in
the Z-scan measurement at each concentration. The linear
relationships between the nonlinear index and the concentrations are shown in Figs. 7 and 8 for two dierent wavelengths. We can see that the value of n2 increases linearly as
the nanocrystal concentration gets higher. The changing
values of the nonlinear refractive index nex indicate that
the nonlinear refraction has contribution not only from
the third-order nonlinearity but also from higher-order
nonlinearities that can be caused by free carrier eect.
The nonlinear refractive index can be written as [17]:
n n0 nex

I n0 n2 n4 II;

4.0

794nm

3.5

3.0

2.5

2.0
0.0

0.4

0.8

1.2

1.6

CdSe Weight Percentage (%)


Fig. 8. Variation between the eective nonlinear refractive index and the
concentration of the CdSe nanoparticles embedded in CR39 at wavelength
794 nm.

where n0 is the linear refractive index, nex is the eective


nonlinear refractive index, I represents the input laser
intensity, n2 is the normal third-order nonlinear refractive
index [19,21], and n4 denotes the fth-order nonlinearity.
Li et al. [21] found that free carrier eect (fth-order nonlinearity) in bulk CdS is signicant for longer laser pulses,
which might still have contribution in our case.
Another interesting eect is due to the local eld eect.
When the nanocrystals (NCs) are embedded in a dielectric
matrix as in the case presented here, the experimentally
measured eective nonlinear optical susceptibility of the
3
composite material veff is related to v3
m through [22]:
3

veff pf 4 v3
m ;

0
is the
where p is the volume fraction of NCs, and f 2e3e0 e
local eld eect (with e0 and e being the dielectric constants
3
of the matrix and nanocrystals, respectively), vm
is the
optical susceptibility of NCs. So in our samples, the change

1.6

397nm

1.4

nex(10 -4 cm /GW)

4.5

nex(10 -4 cm2 /GW)

1444

1.2

of the nonlinear refractive index nex can be caused by the


change of the volume fraction of CdSe nanocrystals in
CR39. It increases when the CdSe nanocrystal weight-percentage concentration increases, which has a linear dependence as shown in Figs. 7 and 8.
4. Conclusions
We have shown that the optical properties of the hybrid
composite lms are strongly dependent on the CdSe nanocrystal concentration in the CR39 polymer. The optical
nonlinearity measurements using the Z-scan technique
have shown that the hybrid materials have eective nonlinear refractive indexes ranging from 2.05 104 cm2/GW to
4.29 104 cm2/GW at 794 nm (below the band gap) and
from 1.09 104 cm2/GW to 1.60 104 cm2/GW at
395 nm (above the band gap), which vary with the input
laser energy as well as the concentration of the CdSe nanoparticles in the polymer. We believe that not only the thirdorder nonlinearity but also higher-order nonlinearity also
exist in such CdSe/CR39 composites due to the free carrier
eect. The dispersion behaviors of the nonlinear refractive
indices at these two wavelengths are consistent with the
bound electronic nonlinear refraction predicted by twophoton resonance. These results indicate that the CdSe/
CR39 lms are a new kind of promising material for optical lter applications and have favorable nonlinear optical
properties.
Acknowledgement
This work was partially supported by National Science
Foundation (NSF SBIR IIP-0538617).

1.0
0.2

0.4

0.6

0.8

1.0

CdSe Weight Percentage (%)


Fig. 7. Variation between the eective nonlinear refractive index and the
concentration of the CdSe nanoparticles embedded in CR39 at wavelength
397 nm.

References
[1] H. Angus Macleod, Thin-Film Optical Filters, Institute of Physics,
Bristol, 2001.

Author's personal copy

C. Gan et al. / Optical Materials 30 (2008) 14401445


[2] H. Yao, S. Takahara, H. Mizuma, T. Kozeki, T. Hayashi, Jpn. J.
Appl. Phys. 35 (1996) 4633.
[3] J. Seo, S. Ma, Q. Yang, L. Creekmore, J. Phys.: Conf. Serie. 38 (2006)
91.
[4] S. Omi, H. Hiraga, K. Uchida, C. Hata, Y. Asahara, A.J. Ikushima,
T. Tokizaki, A. Nakamura: CLE091 Tech. Digest 10 (1991) CtuE2.
[5] R. Adair, L.L. Chase, S.A. Payne, Phys. Rev. B 39 (1989) 3337.
[6] R.E. Schwerzel, K.B. Spahr, J.P. Kurmer, V.E. Wood, J.A. Jenkins,
J. Phys. Chem. A 102 (1998) 5622.
[7] (a) Y. Wang, A. Suna, J. McHugh, E.F. Hilinski, P.A. Lucas, R.D.
Johnson, J. Chem. Phys. 92 (1990) 6927;
(b) L. Qu, X. Peng, J. Am. Chem. Soc. 124 (2002) 2049;
(c) C. Darraud, B. Bennamane, C. Gagnadre, J.L. Decossas, J.C.
Vareille, J. Stejny, Appl. Optics 33 (16) (1994) 3338.
[8] (a) L.E. Brus, J. Chem. Phys. 79 (1983) 5566;
(b) L.E. Brus, J. Chem. Phys. 80 (1984) 4403.
[9] (a) A. Henglein, Chem. Rev. 89 (1989) 1861;
(b) L. Spanhel, M. Haase, H. Weller, A. Henglein, J. Am. Chem. Soc.
109 (1987) 5649.
[10] R. Swanepoel, J. Phys. E: Sci. Instrum. 16 (1983) 1214.

1445

[11] L.W. Wang, A. Zunger, Phys. Rev. B 53 (1996) 9579.


[12] K.C. Krogman, T. Druel, M.K. Sunkara, Nanotechnology 16 (2005)
5338.
[13] K. Akamatsu, S. Deki, J. Mater. Chem. 8 (3) (1998) 637.
[14] X. Wang, Y. Du, S. Ding, Q. Wang, G. Xiong, M. Xie, X. Shen, D.
Pang, J. Phys. Chem. B 110 (4) (2006) 1566.
[15] N. Venkatram, D.N. Rao, Opt. Express 13 (3) (2005) 867.
[16] V.V. Nikesh, A. Dharmadhikari, H. Ono, S. Nozaki, G.R. Kumar, S.
Mahamuni, Appl. Phys. Lett. 84 (23) (2004) 4602.
[17] H. Du, G. Xu, W. Chin, L. Huang, W. Ji, Chem. Mater. 14 (10)
(2002) 4473.
[18] M. Sheik-Bahae, A.A. Said, T.H. Wei, D.J. Hagan, E.W. Van
Stryland, IEEE J. Quantum Electron. QE-26 (1990) 760.
[19] M. Sheik-Bahae, D.C. Hutchings, D.J. Hagan, E.W. Van Stryland,
IEEE J. Quantum Electron. 27 (1991) 1296.
[20] T.D. Krauss, F.W. Wise, Appl. Phys. Lett. 65 (1994) 1739.
[21] H. Li, C.H. Kam, Y. Lam, W. Ji, Opt. Commun. 190 (2001) 351.
[22] J.M. Ballesteros, J. Solis, R. Serna, C.N. Afonso, Appl. Phys. Lett. 74
(1999) 2791.

You might also like