Download as docx, pdf, or txt
Download as docx, pdf, or txt
You are on page 1of 71

SYNTHESIS AND CHARACTERIZATION OF

MONODISPERSE NANOCRYSTALS AND


CLOSE-PACKED NANOCRYSTAL ASSEMBLIES
C. B. Murray and C. R. Kagan
IBM T. J. Watson Research Center, Yorktown Heights, NewYork 10598;
e-mail: cbmurray@us.ibm.com; cheriek@us.ibm.com

M. G. Bawendi
Massachusetts Institute of Technology, Department of Chemistry, Cambridge,
Massachusetts 02139; e-mail: mgb@mit.edu

Key Words quantum dot, nanoparticle, superlattice, colloidal crystal, supercrystal


■ Abstract Solution phase syntheses and size-selective separation methods to pre-
pare semiconductor and metal nanocrystals, tunable in size from ∼ 1 to 20 nm and
monodisperse to ≤5%, are presented. Preparation of monodisperse samples enables
systematic characterization of the structural, electronic, and optical properties of ma-
terials as they evolve from molecular to bulk in the nanometer size range. Sample
uniformity makes it possible to manipulate nanocrystals into close-packed, glassy, and
ordered nanocrystal assemblies (superlattices, colloidal crystals, supercrystals).
Rigorous structural characterization is critical to understanding the electronic and op-
tical properties of both nanocrystals and their assemblies. At inter-particle separa-
tions 5–100 Å, dipole-dipole interactions lead to energy transfer between neighboring
nanocrystals, and electronic tunneling between proximal nanocrystals gives rise to
dark and photoconductivity. At separations <5 Å, exchange interactions cause oth-
erwise insulating assemblies to become semiconducting, metallic, or superconducting
depending on nanocrystal composition. Tailoring the size and composition of the
nanocrystals and the length and electronic structure of the matrix may tune the prop-
erties of nanocrystal solid-state materials.

INTRODUCTION

Many physical phenomena in both organic and inorganic materials have natural
length scales between 1 and 100 nm (102 to 107 atoms). Controlling the physi-
cal size of materials can be used to tune materials properties. In the nanometer
size regime, new mesoscopic phenomena characteristic of this intermediate state
of matter, found in neither bulk nor molecular systems, develop. For example,

0084-6600/00/0801-0545$14.00 545
the electronic and optical properties of metals (1–4) and semiconductors (5–9)
strongly depend on crystallite size in the nanometer size regime. Efforts to ex-
plore structures on the nanometer length scale unite the frontiers of materials
chemistry, physics, and engineering. It is in the design and characterization of
advanced materials that the importance of new interdisciplinary studies may be
realized.
Uncovering and mapping size-dependent materials properties requires synthetic
routes to prepare homologous size series of monodisperse nanometer size crystals,
known as nanocrystals (NCs). NC samples must be monodisperse in terms of size,
shape, internal structure, and surface chemistry. A diverse set of structural probes is
combined to characterize and develop consistent structural models of NC samples.
Optical, electrical, and magnetic studies of well-defined NC samples reveal the
unique size-dependent properties of materials in this intermediate, nanometer size
regime between molecular species and bulk solid.
When atoms or molecules organize into condensed systems, new collective
phenomena develop. Cooperative interactions produce the physical properties we
recognize as characteristic of bulk materials. Like atoms or molecules, but in
the next level of hierarchy, NCs may also be used as the building blocks of con-
densed matter. Routes enabling controlled manipulation of NCs into the glassy
and ordered states of matter lead to the preparation of close-packed NC solids.
Assembling NCs into solids opens up the possibilities of fabricating new solid-
state materials and devices with novel physical properties, as interactions between
proximal NCs give rise to new collective phenomena. Building upon rigorous un-
derstanding of the physical properties of individual NCs, the properties of coupled
NCs in the solids are uncovered. Engineering the size and composition of the
NCs and the length and chemical functionality of the matrix may be used to tune
the unique properties of the individual NC building blocks and those arising from
coupling between proximal NCs.
Although many of the concepts presented in this review are general to a range
of monodisperse NC systems, we often use studies of CdSe NCs to illustrate the
preparation and characterization of NCs and their assemblies (10, 11).

PREPARATION OF MONODISPERSE
NANOCRYSTALS (NCS)
Introduction
The preparation of nearly monodisperse organically passivated NC samples is es-
sential to permit studies that distinguish truly novel properties inherent to nanoscale
structures from those associated with structural heterogeneities or polydispersity.
Although the strict definition of monodisperse requires that particles be identi-
cal or indistinguishable, a relaxed definition is used here (12). Samples with
standard deviations σ ≤ 5% in diameter are referred to as monodisperse. This
corresponds to ± one lattice constant throughout the 1–15-nm size range. NCs
must be uniform not only in size and shape, but they must also have well-formed
crystalline cores and controlled surface chemistry. We reserve the term nanocrys-
tals (NCs) for structures with well-characterized crystalline cores and use the more
general term, nanoparticles, to denote amorphous or inherently multidomain inor-
ganic cores. Although non-crystalline and multidomain nanoparticles can display
a wealth of interesting size-dependent phenomena, this review emphasizes crys-
talline systems. We restrict our discussion to procedures that can reproducibly
prepare a homologous size series of NC samples with rational adjustments of the
experimental conditions and focus on NCs that have been employed as nanoscale
Annu. Rev. Mater. Sci. 2000.30:545-610. Downloaded from www.annualreviews.org

building blocks in assembling new designer solids.


Access provided by University of Pennsylvania on 04/13/16. For personal use only.

General Synthesis and Processing of Monodisperse NCs

Studi klasik oleh La Mer & Dinegar menunjukkan bahwa produksi koloid
monodispersi membutuhkan kejadian nukleasi sementara yang disusul
dengan pertumbuhan yang terkendali secara lambat pada inti yang ada
(Gambar 1A) (13). Penambahan cepat reagen ke bejana reaksi
meningkatkan konsentrasi prekursor di atas ambang nukleasi. Ledakan
nukleasi pendek sebagian mengurangi kejenuhan. Selama konsumsi bahan
baku oleh koloid NCs tumbuh tidak melebihi tingkat

Figure 1 (A) Cartoon depicting the stages of nucleation and growth for the preparation
of monodisperse NCs in the framework of the La Mer model. As NCs grow with time,
a size series of NCs may be isolated by periodically removing aliquots from the reaction
vessel. (B) Representation of the simple synthetic apparatus employed in the preparation
of monodisperse NC samples.
penambahan prekursor untuk solusi, tidak ada bentuk inti baru. Karena
pertumbuhan salah satu NC mirip dengan yang lain, distribusi ukuran awal
sebagian besar ditentukan oleh waktu di mana nukleus terbentuk dan mulai
tumbuh. Jika persentase pertumbuhan NC selama periode nukleasi kecil
dibandingkan dengan pertumbuhan berikutnya, NC dapat menjadi lebih
seragam dari waktu ke waktu (14). Fenomena ini telah disebut sebagai
fokus distribusi ukuran. Banyak sistem menunjukkan fase pertumbuhan
kedua yang berbeda yang disebut pematangan Ostwald (15, 16). Dalam
proses ini, energi permukaan tinggi dari NCs kecil mendorong
pembubarannya, sedangkan material ditempatkan kembali pada NCs yang
Annu. Rev. Mater. Sci. 2000.30:545-610. Downloaded from www.annualreviews.org

lebih besar. Ukuran NC rata-rata meningkat seiring waktu dengan


Access provided by University of Pennsylvania on 04/13/16. For personal use only.

penurunan kompensasi dalam jumlah NC. Pemanfaatan pematangan


Ostwald dapat sangat menyederhanakan persiapan serangkaian ukuran
NCs (17). Bagian-bagian dari campuran reaksi dapat dihilangkan secara
bertahap dalam waktu, seperti yang digambarkan pada Gambar 1A. Bahan
baku sering dapat diekstraksi dengan distribusi awal 10 <σ <diameter 15%,
yang kemudian dipersempit menjadi ≤5% melalui pemrosesan selektif
ukuran.
Preparation of Monodisperse Semiconductor NCs Dalam pertumbuhan senyawa
semikonduktor NCs, supersaturasi yang diperlukan dan nukleasi berikutnya dapat dipicu
oleh injeksi cepat prekursor logam-organik ke dalam labu yang diaduk dengan kuat yang
mengandung pelarut koordinator panas (∼150–350◦C). Pelarut yang biasanya digunakan
adalah campuran alkilfosfin R3P rantai panjang, alkilfosin oksida R3PO (R = butil atau
oktil), alkilamina, dll. (17). Representasi proses sintetik ditunjukkan pada Gambar 1B.
Dalam sintesis II-VI NCs (ME di mana M = Zn, Cd, Hg; E = S, Se, Te), alkil logam
(dimethylcadmium, diethylcadmium, diethylcadmium, diethyl-zinc, dibenzylmercury)
umumnya dipilih sebagai sumber kelompok II. Sumber kelompok VI sering kali adalah
organofosfin kalkogenida (R3PE) atau bistrimetilsilkalkalkogenida TMS2E (TMS,
trimethylsilyl) (di mana E = S, Se, dan Te). Kelas reagen R3PE biasanya lebih disukai
sebagai sumber Se dan Te karena mereka mudah disiapkan; TMS2S dipilih sebagai sumber
S karena lebih reaktif daripada R3PS dan tersedia secara komersial. Penggunaan prekursor
campuran, misalnya kombinasi prekursor Se dan S, mengarah pada produksi langsung
aloy, meskipun stoikiometri NCs tidak secara langsung mencerminkan rasio prekursor,
tetapi lebih pada tingkat diferensial penggabungan prekursor.
Pertumbuhan II-VI NCs tidak terbatas pada penggunaan R3P dan R3PO sebagai pelarut
koordinasi pendidihan tinggi. Injeksi reagen ke dalam alkilfosfat panas, alkilfosfat, piridin,
alkilamin, dan furan semuanya menghasilkan NCs. Mikulec baru-baru ini menunjukkan
bahwa menggunakan prekursor alkylphosphoramide-tellurium, sebagai pengganti R3PTe,
menghasilkan CdTe NCs dengan efisiensi pendaran yang jauh lebih tinggi (18). Interaksi
kuat R3PO dengan Zn prekursor terlalu menghambat pertumbuhan ZnE NCs. Meskipun
R3P masih dapat digunakan, Guyot-Sionnest dan rekan kerja menemukan bahwa
menggunakan alkylamin sebagai pelarut koordinasi sangat meningkatkan tingkat
pertumbuhan ZnE NCs (19).
Demikian pula, sintesis InP dan InAs NCs berkualitas tinggi telah dicapai dengan
mencampur dan memanaskan prekursor III dan V secara cepat dalam pelarut berkoordinasi
tinggi, pelarut terkoordinasi. Persiapan InP dan InAs NCs sekarang mampu menghasilkan
sampel dengan
σ ≤ 10%. Biasanya InCl (C2O4) digunakan sebagai sumber dalam dengan TMS3P atau
TMS3As dalam pelarut R3P / R3PO (20-22). Dalam preparasi III-V ini, precur dalam hadir
dalam pelarut panas sebelum injeksi TMS3P atau TMS3As. Pertumbuhan NCs lambat
karena Ostwald pematangan lebih dari 1 hingga 6 hari diperlukan untuk mencapai ukuran
NC yang diinginkan. Kekayaan prekursor organologam potensial lainnya dan pelarut
koordinasi mendidih tinggi masih belum teruji, sehingga memberikan peluang untuk
ekspansi lanjutan ke sistem NC baru.

Preparation of Monodisperse Metal NCs The synthesis of metal colloids has


been studied for over a century and yet the number of preparations yielding a
size series of monodisperse metal NC samples is surprisingly small. The most
established methods involve aqueous reduction of metal salts (notably Au or Ag) in
the presence of citrate anions (23). These colloids are electrostatically stabilized by
the adsorption of ions to the NCs’ surfaces during growth. These samples have long
been referred to as monodisperse, although in general 10 < σ < 15%. Flocculation
of these colloids is irreversible, preventing further processing to achieve the desired
σ ∼ 5%. Chemisorption of organic ligands on the surface of metal NCs is essential
to permit further handling. Schmid provides an excellent overview of the advances
in metal colloid synthesis (24, 25).
A two-phase reduction method, described by Brust, Schiffrin, and co-workers,
when coupled with size-selective processing produces capped Au and Ag NCs with
σ ∼ 5% (26). In general, aqueous metal salts (e.g. HAuCl4, AgNO3, AgClO4) are
mixed in a toluene solution containing long-chain alkylammonium surfactants to
form a two-phase system. Vigorous stirring for 1 to 3 h transfers the metal salt
into the organic phase, which is then separated. A measured quantity of capping
agent, typically a long-chain thiol, is added to the solution while stirring, and then
a reducing agent (e.g. NaBH4 or hydrazine) is rapidly added to nucleate NCs. The
average NC size is coarsely tunable by adjusting the ratio of capping groups to
metal salt, whereas size-selective precipitation is employed to narrow the initial
size distribution. Several studies have refined the preparation of thiol capped Ag
and Au NCs (27, 28).
The preparation of metal NCs in inverse micelles warrants mention. The inverse
micelle method has been employed since the late 1980s for the preparation of both
semiconductor and metal NCs. Although it is widely adopted, samples approach-
ing the desired σ ≤ 5% are rarely observed. However, Pileni and co-workers
provide a notable exception by coupling the initial synthesis with extensive use
of size-selective precipitation to yield high quality Ag (29), AgS (30), and more
recently Co NCs (31).
Higher temperature reduction of metal salts in the presence of stabilizing agents
can also be employed to produce monodisperse transition metal (e.g. Co and
Ni) NCs that do not crystallize well at room temperature (RT) (32). In this general
scheme metal halides or acetates are dissolved in high-boiling inert sol- vents
(e.g. octylether, phenylether) along with a combination of R3P and long- chain
carboxylic acids (e.g. oleic acid). The solution of metal salts and stabi- lizers is
vigorously stirred and heated to ∼ 200–250◦ C at which time a solution
containing a strong reducing agent [e.g. LiHB(CH2CH3)3, Na naphthalide, etc] is
injected. Metal NCs nucleate and grow until the reagent is consumed. Although no
Ostwald ripening is observed, NC size is coarsely tunable by the ratio of capping
groups to metal salt. Size-selective precipitation yields NC samples with σ ∼ 5%.
Progress has also been made in the preparation of monodisperse bimetallic NCs.
For example, see the work by Bradley and co-workers (33). Figure 2A–D shows
high-resolution transmission electron microscopy (HRTEM) images of some of
the NC materials that can currently be prepared and isolated.

NC Core/Shell Structures Techniques to overcoat NCs with an inorganic shell


Annu. Rev. Mater. Sci. 2000.30:545-610. Downloaded from www.annualreviews.org

are remarkably general with only a few modest constraints: (a) The existing NC
Access provided by University of Pennsylvania on 04/13/16. For personal use only.

Figure 2 Collection of high resolution TEM images for typical NC materials such as (A)
h100i-oriented CdSe (scale bar = 15 Å), (B) h001i-oriented CdSe (scale bar = 15 Å),
(C) CdTe (scale bar = 20 Å), and (D) Co (scale bar = 25 Å) (31) [CdSe images courtesy
of Kadavanich (62); CdTe courtesy of F Mikulec].
seeds must withstand the conditions under which the second phase is deposited,
(b) the surface energies of the two phases must be sufficiently similar so that the
barrier for heterogeneous nucleation of the second phase is lower than that for
homogeneous nucleation, and (c) the seed NC and the overcoat material must
not readily interdiffuse under the deposition conditions. Typically seed NCs are
prepared and isolated by one of the standard procedures outlined above, size-
selected, and then redispersed in a fresh solution of solvent and stabilizers. The
solution is then heated while precursors for the inorganic shell are gradually added
to allow the material to heterogeneously nucleate on the seed NCs. If the rate of
precursor addition does not exceed the rate of deposition on the seeds, the precursor
Annu. Rev. Mater. Sci. 2000.30:545-610. Downloaded from www.annualreviews.org

concentration never reaches the threshold for homogeneous nucleation of a second


Access provided by University of Pennsylvania on 04/13/16. For personal use only.

inorganic phase.
Methods for overcoating a semiconductor NC with a second semiconductor
material of wider bandgap are well developed. For example, CdSe nanocrystals
have been overcoated with ZnS (34–36), ZnSe (37), and CdS (38), which resulted
in dramatic improvements in luminescence efficiency, exemplified by the work on
CdSe/ZnS by Guyot-Sionnest and co-workers (35). Mews, Weller, and co-workers
have overcoated CdS NCs with lower bandgap materials such as HgS (39), while
others have coated CdTe cores with HgTe shells (40). The HgS shell in CdS/HgS
systems has in turn been buried by the deposition of new CdS material to produce
a shell of low-bandgap HgS nested in a wider bandgap NC structure (41).
Peng and co-workers showed that steady secondary addition of reagents, com-
mon for overcoating NCs, could also dramatically improve the synthesis of single
component systems (42). By following the initial nucleation of II-VI and III-V
NCs with a calculated secondary addition of the original precursors, NC growth
is accelerated and the size distribution is focused, as predicted by Reiss (14). This
modified synthesis often yields samples where 5% < σ < 10% in the reaction
flask, thus minimizing the need for size-selective processing.

Cap (Ligand) Exchange


Stabilizing agents must be present during growth to prevent aggregation and pre-
cipitation of the NCs. When the stabilizing molecules are attached to the NC
surface as a monolayer through covalent, dative, or ionic bonds, they are referred
to as capping groups (43). These capping groups serve to mediate NC growth,
sterically stabilize NCs in solution, and passivate surface electronic states in semi-
conductor NCs. This surface capping is analogous to the binding of ligands in
more traditional coordination chemistry. Synthetic organic techniques allow the
tail and head groups to be independently tailored through well-established chem-
ical substitutions. NC surface derivatization can be modified by ligand exchange:
repeated exposure of the NCs to an excess of a competing capping group, followed
by precipitation to isolate the partially exchanged NCs (10, 17, 44). Repeating this
cycle allows more complete exchange. This recursive approach can cap NCs with
a wide range of chemical functionalities, even if the binding of the new cap is less
favorable than the original. The cap exchange process has been used extensively
to adjust the dimensions of the organic layer surrounding the NCs and thus the
minimum inter-particle spacing in NC assemblies (45).

Isolation and Purification of NCs


NCs are stable with respect to aggregation only if the capping groups provide a
repulsive force of sufficient strength and range to counteract the inherent van der
Waals attraction between NCs. The energetic barrier to aggregation provided by
the capping groups is strongly dependent on the energy of mixing between the
Annu. Rev. Mater. Sci. 2000.30:545-610. Downloaded from www.annualreviews.org

tethered capping groups and the solvent. Introduction of a nonsolvent, miscible


Access provided by University of Pennsylvania on 04/13/16. For personal use only.

with the original dispersing solvent, destabilizes the NC dispersions. The NCs
then aggregate and precipitate leaving many of the synthetic by-products in solu-
tion. If the capping groups are well bound to the surface of the NCs, the resulting
powders are redispersible in a variety of solvents: alkanes, aromatics, long-chain
alcohols, chlorinated solvents, and organic bases (e.g. amines, pyridines, furans,
phosphines). Repeated flocculation and redispersion of the NCs in fresh solvents
allow the isolation of powders composed of the desired NCs and their intimate or-
ganic capping layer (17). A straightforward extension of this precipitation process
allows the isolation of size-selected fractions of NCs (46).
Size-selective precipitation involves either the titration of a nonsolvent into
the dispersion (17) or the preferential evaporation of solvent from a mixed sol-
vent/nonsolvent system to bring about gradual flocculation (47). Since the largest
NCs in the size distribution exhibit the greatest attractive van der Waals forces,
they tend to aggregate before the smaller NCs. As aggregates of larger NCs
form, there is a natural tendency to exclude smaller NCs (48, 49). If the disper-
sion is allowed to only partially flocculate, filtering or centrifuging the suspension
isolates a precipitate enriched in the larger NCs and leaves the smaller NCs dis-
persed in the supernatant. The precipitate can be redispersed in a solvent and
subjected recursively to this gentle flocculation procedure to further narrow the
size distribution. Similarly, gradual addition of more nonsolvent to the decanted
supernatant brings about a precipitation of a second size fraction. Size-selective
precipitation is analogous to purification by fractional crystallization (50) or long-
standing methods to fractionate polymers on the basis of molecular weight. Nar-
rower initial size distributions allow the desired σ value to be attained with fewer
stages of size-selective precipitation and thus higher yield. Slower destabilization
leads to more efficient separation of the sizes, yet even well-practiced separations
of NCs with an initial σ ∼ 10%, yields ≤30% of the initial NC sample with
σ ∼ 5%.
Gel permeation (size exclusion) chromatography can also be employed to nar-
row an initially broad size distribution, although once again the yield is limited and
the process is slow (51). Wilcoxon and co-workers provide some current examples
of efforts to separate NCs using chromatographic techniques (52).
CHARACTERIZATION OF THE NC CORE AND SURFACE

Introduction
Many studies have investigated the structure and chemical composition of NCs in
an effort to reveal the structure/property relationships as the crystals grow from
molecular species toward bulk solids. This is a tremendous interdisciplinary chal-
lenge because the smallest NCs (<1 nm) are nearly molecular (<100 atoms) and
the largest NCs (>20 nm) contain >100,000 atoms. In this size range the percent-
age of surface atoms goes from >75% to <0.5%. Standard chemical and surface
sensitive probes are well suited for studies of the smallest species whereas physical
Annu. Rev. Mater. Sci. 2000.30:545-610. Downloaded from www.annualreviews.org
Access provided by University of Pennsylvania on 04/13/16. For personal use only.

probes, which exploit the periodicity of the internal NC lattice, are better suited
for the largest species.

Chemical Analysis of NCs


Detailed elemental analysis is the first essential step in developing a model of NC
structure. Techniques as diverse as atomic absorption and emission, neutron acti-
vation, X-ray fluorescence, and mass spectroscopy (53) can be employed to reveal
the composition of the average NC. The use of specific chemical spectroscopies,
including infrared absorption and solution phase and H1 and C13 nuclear magnetic
resonance (NMR) (54–55), photoelectron spectroscopy (56), and electron energy
loss spectroscopy (EELS) (57), often aid in determining not only what species are
present but how the components are distributed between NC surface and core. In
addition, solid-state Se77 (58) and P31 (59) NMR have been used to investigate the
inorganic species at the NC/organic interface of CdSe and InP, respectively. When
these studies are combined, rational descriptions of the organic capping layer and
the near surface region of the NCs emerge.
Structural probes such as transmission electron microscopy (TEM), extended
X-ray absorption fine structure (EXAFS) spectroscopy, small-angle X-ray scat-
tering (SAXS), and wide-angle X-ray scattering (WAXS) are powerful probes of
NC shape and internal structure. If the results of these diverse techniques can be
made consistent with a single atomistic model of average NC structure, a coherent
description of the system is possible.

Transmission Electron Microscopy of NCs


TEM is arguably the single most powerful technique for characterizing NCs and
their assemblies. Routine low-resolution TEM studies image ensembles of NCs,
permitting a statistical description of the size and shape of NCs in a sample to
be developed. HRTEM imaging reveals the individual NC shape and internal
structure. TEM studies of small metal NCs have been reviewed (60, 61). The
methods developed for metal NCs are equally applicable to the study of dielectric
and semiconducting NCs.
Examples of HRTEM images for CdSe (62), CdTe (18), and Co NCs (32) are
seen in Figure 2. The relationship between the observed interference fringes and
the underlying crystal lattice can be complex, and care must be taken when using
real-space images as evidence for subtle lattice reconstructions or precise measures
of lattice constants (63, 64). However, it is useful to develop a list of the observed
structural motifs and their relative probabilities. Evidence of new structural phases
of metals may also be observed in TEM of small NCs. Figure 2D shows an image
of a well-formed 80 Å diameter Co NC with a cubic lattice structure that does not
correspond to the known hcp or fcc phases of cobalt. The NC structure is that of
the beta phase of manganese and the label ε-Co has been proposed (65) for this
Annu. Rev. Mater. Sci. 2000.30:545-610. Downloaded from www.annualreviews.org

new form of Co metal. A number of studies also indicate that multiply twinned
Access provided by University of Pennsylvania on 04/13/16. For personal use only.

structures may in fact be the lowest energy configurations for many metal NC
systems (60, 61).
These rough statistics are used to form an initial set of parameters for simulations
of NC structure. Extensive surveys of NC samples ranging from ∼25 to 150 Å in
diameter yield some general qualitative and quantitative results regarding NC size,
shape, and structure. In an effort to provide accurate measurements of size, the
lattice fringes of the NCs are used as an internal standard. Counting the number
of lattice planes or atom columns in each NC is used to determine NC size and
calibrate the magnification of the image.
Measurement of several hundred NCs in a series of images, like those shown
in Figure 3, are employed to develop a statistical model of NC size and shape,
while 10 to 20 HRTEM images, of the type shown in Figure 2A–D, are used to
develop a description of internal NC structure. These limited observations must be
compared with X-ray scattering studies, which simultaneously probe statistically
large ensembles of NCs.

Small-Angle X-Ray Scattering of NCs


SAXS studies of monodisperse NC samples reveal scattering structure previously
cloaked by polydispersity and have greatly simplified the analysis (10, 66–69). The
SAXS intensity, I(q), scattered from a collection of N, non-interacting particles,
of rigorously uniform electron density, ρ, in a homogeneous medium of density,
ρ o, is given by
I(q) = Io N(ρ − ρo )2 F2 (q). 1.
F(q) is the form factor—the Fourier transform of the shape of the scattering object.
For spheres of radius R, the form factor is expressed by (70, 71)
· ¸
4 3 sin(qR) −qR cos(qR)
F(q) = πR 3 . 2.
3 (qR)3

Figure 4A curve (a) shows the calculated pattern for a collection of 62 Å monodis-
perse spheres of rigorously uniform electron density. Real objects are not of
uniform electron density. Each atom in the structure provides a local modulation
Annu. Rev. Mater. Sci. 2000.30:545-610. Downloaded from www.annualreviews.org
Access provided by University of Pennsylvania on 04/13/16. For personal use only.

Figure 3 Large-field TEM images are employed to develop statistics on NC size and
shape. A collection of 48 Å CdSe NCs at (A) low magnification (scale bar = 200 Å) and
(B) higher magnification (scale bar = 80 Å) (45); monolayer of 80 Å Co NCs at (C ) low
magnification (scale bar = 500 Å) at (D) higher magnification (scale bar = 65 Å) (32).

that introduces a broad scattering background. Curve (b) shows the scattering
pattern for a 4500-atom spherical fragment (∼62 Å) of the bulk CdSe wurtzite
lattice. The signal is the sum of the scattering from atomic sites (the broad back-
ground) as well as the coherent scattering from all discrete distances within the
NC. Scattering from atomic sites scales like the volume (∝R3), while the inten-
sity of the oscillations from NC shape scales like the square of the volume (∝R6).
In large particles, such as micron size latex and silica, the contribution from atomic
scattering becomes negligible. Neglecting this contribution in NC systems,
Annu. Rev. Mater. Sci. 2000.30:545-610. Downloaded from www.annualreviews.org
Access provided by University of Pennsylvania on 04/13/16. For personal use only.

Figure 4 (A) SAXS patterns for model structures having 4500 atoms, comparable to a
62 Å diameter CdSe NC (symbols). The curves are models for (a) 62 Å spheres of uniform
electron density; (b) monodisperse, 4500 atom spherical fragments of the bulk CdSe lattice;
(c) monodisperse, 4500 atom ellipsoidal fragments of the bulk CdSe lattice, having a 1.2
aspect ratio; and (d) fit to SAXS data (dots) assuming a Gaussian distribution of ellipsoids
(as in curve c), yielding the NC sample size and size distribution. (B) SAXS patterns for
CdSe NC samples ranging from 30 to 75 Å in diameter (dots). Fits are used to devise the
NC sample size, reported in equivalent diameters, and size distributions, ranging from 3.5
to 4.5% for the samples shown.
however, would result in an overestimation of the size distribution. TEM stud-
ies indicate that many NC systems are better described as ellipsoids than spheres
(17, 62). The scattering from elliptical particles has been dealt with extensively
elsewhere (72, 73). We use an atomistic model, discussed below, to calculate
SAXS scattering. The model system is essentially a collection of ellipsoidal frag-
ments of the bulk lattice of the respective NCs. Although the detail of the internal
structure has little effect of the SAXS pattern, the atomistic model also reproduces
the wide-angle scattering as well. Small variations in NC size and shape quickly
wash out the oscillations in SAXS. The amplitude of the oscillations allows deter-
mination of NC size and size distribution with greater confidence than any other
Annu. Rev. Mater. Sci. 2000.30:545-610. Downloaded from www.annualreviews.org

structural technique. Curve (c) shows the predicted pattern for a monodisperse
Access provided by University of Pennsylvania on 04/13/16. For personal use only.

ensemble of prolate (aspect ratio 1.2 as determined by TEM) fragments of the bulk
CdSe lattice containing 4500 atoms. Curve (d) (dots) shows the experimental re-
sults and fit (solid line) to an atomistic model of SAXS scattering created from the
statistical observations of TEM. Experimental SAXS patterns (dots) and computer
simulations (solid lines) are used to measure size and size distribution for a size
series of CdSe NCs (Figure 4B). Korgel & Fitzmaurice have recently published a
detailed study of monodisperse Ag NCs using SAXS (68, 69).

Wide-Angle X-Ray Scattering of NCs


WAXS like SAXS probes a statistically large number of NCs. WAXS reveals the
internal structure of the average NC core and permits measurement of NC size
and shape (17, 74, 75). Experimental WAXS patterns for CdSe NCs ranging from
17 to 90 Å in diameter (Figure 5A) show evidence of finite size broadening in all
reflections. The diffraction feature centered at ∼ 25◦ 2θ is a convolution of several
crystallographic reflections, and thus its average position and width are not appro-
priate for simple estimates of lattice parameters or NC size. Excessive attenuation
and broadening of (102) and (103) reflections are signatures of stacking faults
along the h002i axis, as observed in HRTEM. Experimental and computational
WAXS studies of NC ensembles demonstrate the sensitivity of diffraction to NC
size, shape, and planar disorder and to thermal effects (17, 74). These studies
are extended to probe the evolution of NC structure with size and address the
importance of NC shape and the possibility of lattice contractions.
The intensity of X-ray scattering, I(q), is described by the Debye equation:
XX sin(qrmn )
I(q) = I0 Fm Fn , 3.
qrmn
m n
where Io is the incident intensity, q = 4π sin(θ )/2λ is the scattering parameter for
X-rays of wavelength λ diffracted through the angle θ, rmn is the distance between
atoms m and n, with atomic form factors Fm and Fn, respectively (75). A discrete
form of the Debye equation is given by
f 2 (q) X ρ(rk )
I(q) = Io sin(qr), 4.
q rk
k
Annu. Rev. Mater. Sci. 2000.30:545-610. Downloaded from www.annualreviews.org
Access provided by University of Pennsylvania on 04/13/16. For personal use only.

Figure 5 (A) WAXS patterns for CdSe NC samples ranging from 17 to 90 Å in di-

ameter. (B) Three-dimensional representation of a CdSe NC, as developed from TEM


studies, exemplifies the atomistic structure employed in SAXS and WAXS modeling (62).
(C ) SAXS and WAXS pattern for an ∼4500 atom (62 Å) CdSe NC samples (dots). Simul-
taneous fitting, using nonlinear least squares methods, to the SAXS and WAXS patterns
fits the sample average NC size to ∼4500 atoms, aspect ratio to 1.2 (prolate), and size
distribution to 4.2%.

where the summation is taken over all inter-atomic distances rk, which occur with
the number ρ(rk) times in a given NC (75). Since the number of unique inter-
atomic distances in an ordered structure grows much more slowly than the total
number of distances, using the discrete form of the equation is significantly more
efficient in simulating scattering from large NCs.
Atomic coordinates for the simulated NCs are obtained by systematically gen-
erating positions for atoms in a bulk crystalline lattice falling within a defined
ellipsoid. Approximate NC size and shape measured by TEM are used to initially
define the dimensions of the model ellipsoids that are then refined in simulations
by simultaneous fitting of the WAXS and SAXS patterns (10). In the case of
CdSe, planar disorder along the h002i axis is extracted by fitting WAXS patterns
to atomic coordinates for a set of model structures having n randomly distributed
stacking faults, where n = 1, 2, or 3. Thermal effects are simulated by introducing
a Debye-Waller factor. A schematic of such an atomistic model is shown in Figure
5B. Figure 5C shows a representative simultaneous fit of a SAXS pattern, collected
from NCs dispersed at 1% by weight in polymer, and a WAXS pattern, collected
from a NC powder, of the same 64 Å CdSe NC sample with ∼ 4.2% size distribution.
Simultaneous fitting of the SAXS and WAXS patterns using parameters con-
sistent with TEM observations and EXAFS studies (discussed below) unites the
results of these diverse structural probes to yield a description of average NC struc-
ture. The small NCs are often referred to as clusters because they possess too few
atoms to define a core crystal structure and thus distinguish between structural
motfis. This is observed in diffraction patterns of the smallest members of the
Annu. Rev. Mater. Sci. 2000.30:545-610. Downloaded from www.annualreviews.org

CdSe series (Figure 5A), 17 to 22 Å in diameter, as an abrupt transition in the


Access provided by University of Pennsylvania on 04/13/16. For personal use only.

WAXS patterns. This transition in the diffraction features indicates a significant


surface reconstruction or change in NC shape. In 17 Å NCs, all the atoms in the
cluster are less than two bond lengths from the surface, and thus a local structural
probe of bond length, such as EXAFS, is better suited to their characterization.
EXAFS refers to the complex, oscillatory character of the X-ray absorption of
a material at energies between 40 and 1000 eV above its characteristic absorp-
tion edge. This oscillatory behavior depends on the environment of the absorbing
species and provides a measurement of bond length (76). Studies, for example
on 15 and 35 Å CdE (E = S, Se, and Te) NCs (77), can be fit with atomistic

models of NC structure and are consistent with NC structural models developed


from WAXS, SAXS, and TEM measurements. EXAFS has also been used to study
the local structure of bimetallic NCs (78).
The smallest species of CdS have been isolated as molecular crystals, permit-
ting unambiguous determination of the internal NC structure using single-crystal
diffraction techniques (79, 80). In an elegant series of experiments, Weller and co-
workers made direct comparisons between EXAFS and single-crystal diffraction
results (81). The application of a diverse array of common chemical and structural
probes to a homologous size series of monodisperse samples yields self-consistent
models of the average NC structure in many materials systems.

OPTICAL AND ELECTRONIC PROPERTIES


OF DISPERSED NANOCRYSTALS
Optical Properties
Nanometer size semiconductor NCs smaller than the Bohr radius of the bulk
exciton strongly confine electronic excitations in all three dimensions (82, 83).
Photoexcitation of the NC creates an electron-hole pair that is confined to and
delocalized over the volume of the NC. The photophysical properties of the NC
are analogous to those of a large molecule. In the strong confinement regime, the
electron and hole can be treated independently, and the electronic structure of the
NC can be modeled using simple effective mass theory (see the review by Efros
& Rosen, this volume). The effective mass approximation assumes parabolic con-
duction and valence bands with bulk effective masses for the electron and hole.
Each carrier can be treated as a particle in a sphere bound at the NC surface by
an infinite potential. The electron and hole in the NC are described by hydro-
genic wavefunctions and occupy discrete electronic energy levels. The NC has
been coined the artificial atom because of the atomic-like nature of its electronic
wavefunctions and energy levels.
Three-dimensional confinement effects collapse the continuous density of states
of the bulk solid into the discrete electronic states of the NC. The finite size of the
NC quantizes the allowed k values (Figure 6 left). Decreasing NC diameter shifts
Annu. Rev. Mater. Sci. 2000.30:545-610. Downloaded from www.annualreviews.org

the first state to larger k values and increases the separation between states. This is
Access provided by University of Pennsylvania on 04/13/16. For personal use only.

observed spectroscopically (Figure 6A) for a series of CdSe NC samples, as a blue

Figure 6 The bulk conduction and valence bands for semiconductors are assumed to be
parabolic in the simple effective mass approximation. Energy diagrams (E versus k) show
the complexity of the valence band for the example of CdSe, important in assigning NC
electronic states. The finite size of the NC quantizes the allowed k values. Decreasing
the NC diameter shifts the first state to larger values of k and increases the separation
between states. (a) This is seen spectroscopically as a blue shift in the absorption edge
and a larger separation between electronic transitions for a homologous size series of CdSe
NC dispersions, collected at RT. (b) Observation of discrete electronic transitions in optical
absorption is a measure of the wealth of spectroscopic information that can be uncovered
in monodisperse NC samples (σ ≤ 5%).
shift in the absorption edge and a larger separation between electronic transitions
with decreasing NC diameter (10, 17).
The preparation of monodisperse NC samples has made it possible to observe,
assign, and monitor the size evolution of a series of discrete, excited electronic
states (Figure 6B). These samples have strong band-edge emission with photo-
luminescence (PL) quantum yields (QYs) ranging from 0.1 to 0.9 at 10 K (84).
Overcoating the core semiconductor NC with a larger bandgap semiconductor
does a better job at electronically passivating surface states and has made QYs of
≥0.5 at RT routine (38). Norris extracted the size-dependent optical absorption
and emission spectra of single CdSe NCs using transient differential absorption,
Annu. Rev. Mater. Sci. 2000.30:545-610. Downloaded from www.annualreviews.org

photoluminescence excitation (PLE) (85), and fluorescence line narrowing (FLN)


Access provided by University of Pennsylvania on 04/13/16. For personal use only.

spectroscopies (84). Since then other semiconductor NC samples, such as InAs and
InP, have similarly been characterized (86, 87). Agreement between experimental
observations and theoretical calculations provides a framework for understanding
the size-dependent optical spectrum of individual NCs. PL studies of single CdSe
NCs reveal ultra-narrow homogeneous linewidths for the NC emission, signifi-
cantly narrower than those obtained previously using FLN (88). The electronic
structure of the NC is truly atomic-like (89).
As the size of metal NCs decreases to length scales smaller than the wavelength
of incident radiation, a surface plasmon resonance develops—for noble metals it
lies within the visible (90). The surface plasmon resonance is a dipolar excitation
of the entire particle between the negatively charged free electrons in the NC and its
positively charged core. The energy of the surface plasmon resonance depends on
both the free electron density and the dielectric medium surrounding the NC (see
Exchange Coupling). The width of the resonance varies with the characteristic
time before electron scattering. For larger NCs, the resonance sharpens as the
scattering length, defined by the distance to the NC surface, increases.
If the size of metal NCs is made small enough, the continuous density of
electronic states is broken up into discrete energy levels. The spacing, δ, between
energy levels depends on the Fermi energy of the metal, EF, and on the number
4EF
of electrons in the metal, N as δ = 3N (1). The discrete electronic energy levels
in metal NCs has recently been observed in far-infrared absorption measurements
of Au NC samples (53). At finite size, the evolution of properties for metals from
the atomic level to bulk solid is observable.

Electronic Properties
As the size of semiconductors and metals approach nanometer length scales, the
finite size of the materials also leads to unique electronic properties. To add a sin-
gle charge to a semiconductor or metal NC costs energy because a charge carrier
is no longer solvated in an effectively infinite medium. For a NC surrounded by
a medium of dielectric constant, ε, the capacitance of the NC depends on its size
as C(r) = 4π ε orε and the energy required to add a single charge is given by the
e2
charging energy EC = 2C(r) (7, 91). Tunneling of single charges onto metal or
Annu. Rev. Mater. Sci. 2000.30:545-610. Downloaded from www.annualreviews.org
Access provided by University of Pennsylvania on 04/13/16. For personal use only.

Figure 7 (a) I-V curves show the effects of finite size on conduction for a single 10 nm
Al NC. At low temperature, the energy cost to add charge to a NC gives rise to an onset to
current flow, Coulomb blockade. (b) For small enough metal NCs, the density of electronic
states are quantized. The I-V curve shows steps just above the Coulomb blockade as current
arises from tunneling through the discrete energy levels of the NC (courtesy of CT Black).

semiconductor NCs can be seen at temperatures, kBT < EC, in the I-V character-
istics from devices containing single NCs (92, 93) or from STM measurements of
NCs on conductive surfaces (94, 95). Figure 7 shows, for a 100 Å Al NC, that the
charging energy gives rise to a barrier to current flow known as Coulomb blockade
(96). Adding a gate electrode to the structure can be used to modulate the chemical
potential of the NC and therefore the voltage for current flow. This three-terminal
device is known as a single-electron transistor and has received great attention as
an exploratory device structure (97, 98).
At low temperatures and for small metal NCs, for example ∼1 K and 120 Å for
Al, quantization of the NC electronic energy levels is also observable in electrical
measurements. Figure 7b, a magnification of Figure 7a, shows structure in the I-V
characteristics above the threshold for carrier tunneling. These steps in the I-V
characteristics correspond to tunneling via the discrete energy levels in the NC.
For NCs small enough to show Coulomb blockade, but too large to demonstrate
the discrete electronic energy levels, the I-V characteristics show a continuous rise
above the onset of current flow.

SELF-ORGANIZATION OF CLOSE-PACKED NC SOLIDS

Introduction
Organization of uniform particles into ordered assemblies occurs in many natural
systems. Striking examples include the geologic formation of opals (99), whose
prized iridescence results from Bragg diffraction of light by the regular lattice
Annu. Rev. Mater. Sci. 2000.30:545-610. Downloaded from www.annualreviews.org
Access provided by University of Pennsylvania on 04/13/16. For personal use only.

of silica particles from which it is composed, and the biological formation of virus
crystals (100). The assembly of synthetic colloidal particles 0.1–1 µm in
diameter is under active study because these and related artificial systems are in-
teresting photonic materials for their potential, yet unrealized “photonic bandgap”
(101, 102). In each case, it is the combination of particle uniformity and carefully
controlled growth environment that permits the formation of these colloidal solids.
Precise control over NC size, shape, composition, and surface chemistry permits
rational assembly of NCs into close-packed solids. Small CdS clusters, containing
<100 atoms, are truly monodisperse and have been isolated as true molecular crys-
tals (79–81). Ordering of amorphous iron oxide particles, observed by Bentzon
et al, is one of the earliest examples of artificial nanoparticle assemblies (103, 104).
Controlling electronic coupling between semiconductor and metal NCs can now
be explored when extremely strict tolerances both in the preparation and character-
ization of individual NCs and their assemblies are met. The spacing between NCs
in the solids can be varied from intimate contact to ∼50 Å. Glassy NC solids, with
only short-range order and random NC orientation, provide isotropic materials,
whereas three-dimensional NC superlattices, ordered over hundreds of microns
with preferred NC orientations, produce highly anisotropic media (45).
Methods have been developed for the preparation of amorphous (glassy) and
crystalline arrays of NCs both as freestanding structures and as epitaxial thin
films. Systematic study of both disordered and ordered phases is essential to dif-
ferentiate the contributions of NC proximity and periodicity to the development
of new collective physical phenomena.

Preparation of Locally Ordered NC Assemblies


(Colloidal Glasses)
Close-packed glassy solids are prepared by tailoring the solvent composition to
maintain stability of the NC dispersion as it is concentrated by solvent evapo-
ration (10, 11, 66, 67, 105, 106). When the interaction between NCs is weak and
repulsive, there is no significant energy driving the formation of an ordered lattice.
As the concentration rises with solvent evaporation, the free volume available to
each NC is smoothly reduced, and the viscosity of the dispersion increases until it
freezes the local liquid-like structure in the solid (107–109). The formation of NC
glasses is reversible at any step by gradual addition of fresh solvent. These solids
have liquid-like radial distribution functions, resembling the random packing of
hard spheres with soft shells. Measurements of the glass density exceed 80% space
filling and are consistent with the dense random packing of hard spheres with soft
shells (capping groups filling the interstices). This random close packing is evident
in the HRSEM image (Figure 8a) of 56 Å CdSe NCs. At higher magnification (in-
set), the micrographs reveal that the NCs are close-packed, forming a glassy solid
in which each NC remains separated from its neighbors by the organic capping
groups. The glasses are brittle materials that fracture under stress. Figure 8b shows
Annu. Rev. Mater. Sci. 2000.30:545-610. Downloaded from www.annualreviews.org

the fracture surface of a ∼240-µm thick NC glass prepared by slow evaporation


Access provided by University of Pennsylvania on 04/13/16. For personal use only.

of a dispersion of 38 Å NCs. The conchoidal fracture lines are characteristic of


brittle failure in glassy materials. Neither macroscopic nor microscopic porosity
is evident.
SAXS studies allow the average structure in these NC glasses to be modeled.
Each NC acts as a scattering center with a form factor characteristic of its size
and shape (see Small-Angle X-Ray Scattering of NC). In NC glasses, there is
additional contribution to the SAXS. In concentrated dispersions and colloidal
and NC solids, the positions of the particles become correlated, and the SAXS
is modulated by the interference between particles (structure factor) (110–112).
Figure 9A shows SAXS spectra, plotted on a log scale, for a size series of NC
samples, dispersed at ∼1% in poly-vinylbutyral (PVB) (dotted lines) and close-
packed into ∼1 to 0.5 µm thick glassy solids (solid lines). It is important to note
that although the inter-particle interference produces a peak in the SAXS pattern,
direct application of the Bragg equation is not appropriate. Strong contributions
from both the structure factor of the solid and the form factors of the individual
NCs influence peak position, and therefore the peak in the SAXS pattern does not
directly correspond to inter-particle spacing.
The scattered intensity, I(q)total, from an arbitrary collection of N particles is
I(q)total = Io NF2 (q)S(q), 5.
where the structure factor S(q) is the interference introduced by the correlation
between particle positions. The structure factor can be extracted from the exper-
imental data by dividing the SAXS pattern of the NC glasses by that of the NCs
dispersed in PVB. When the NCs are dilute, their positions are uncorrelated and
S(q) ∼ 1. Equation 6 expresses the structure factor in terms of an average radial
distribution ρ(r) of particles about an arbitrary reference particle (110).
Z · ¸
4πρo ∞ ρ(r)
S(q) = 1 + r − 1 sin(qr) dr 6.
q 0 ρo

The structure factor can be Fourier transformed to yield a pair distribution function
(PDF) g(r) where
Z∞
ρ(r) 1
g(r) = =1+ [S(q) − 1]q sin(qr) dq. 7.
ρo 2π 2 rρ o 0
Annu. Rev. Mater. Sci. 2000.30:545-610. Downloaded from www.annualreviews.org
Access provided by University of Pennsylvania on 04/13/16. For personal use only.

Figure 8 (a) Low and high (inset) magnification HRSEM micrographs of a glassy NC
solid prepared from 56 Å CdSe NCs. The NCs are randomly close-packed in the solid with
each NC separated from its neighbors by the organic cap. (67). (b) SEM micrograph of the
fracture surface from a ∼240-µm-thick glassy NC solid of 38 Å CdSe QDs. The NC solid
was imaged by tilting it 45◦ with respect to the incident electron beam (66).
Annu. Rev. Mater. Sci. 2000.30:545-610. Downloaded from www.annualreviews.org
Access provided by University of Pennsylvania on 04/13/16. For personal use only.

Figure 9 (A) Comparison of experimental SAXS patterns for CdSe NC samples, ranging
from 32 to 72 Å in diameter, dispersed in PVB (dotted lines) and close-packed into glassy
solids (solid line). (B) Pair distribution functions (PDFs), g(r) extracted from experimental
SAXS data. TOPO/TOPSe caps maintain an average inter-particle spacing of 11 ± 1 Å.
Annu. Rev. Mater. Sci. 2000.30:545-610. Downloaded from www.annualreviews.org
Access provided by University of Pennsylvania on 04/13/16. For personal use only.

Figure 10 Cartoon depicting the local arrangement of NCs in both glassy and ordered
QD solids. SAXS measurements on monodisperse NC samples: those that are dispersed
in a polymer yield the radius (R) of the NCs; those that are close-packed into NC solids
give the nearest neighbor distance (NN). A combination of these measurements yields the
inter-particle separation (d) between NCs in the solids. Cap-exchange may be used to tune
the inter-particle separation and the electronic structure of the organic inter-layer.

Experimental observations are limited to 2θ angles between 1◦ and 12◦ , and thus
the data must be rationally extrapolated at the ends of the experimental range to
minimize truncation errors (113). Figure 9B shows the corresponding PDFs for
each of the NC glasses, the first peak representing the average center-to-center
distance between NCs in the solids. Combining SAXS measurements of NC
size and size distribution from dispersed samples, with measurements of nearest
neighbor distance from NC glasses, defines all the length scales in the NC glass
(Figure 10). For a R3PO/R3P (R = octyl) capping layer, the nearest neighbor
distance, d, between NCs is 11 ± 1 Å.

Engineering the Spacing Between NCs


Cap exchange techniques are used to provide precise control over NC spac-
ing in the solids (10, 45). SAXS shows (Figure 11A, B) that when the native
R3PO/R3P capping groups R = octyl [curve(a)] are exchanged for the less bulky
R = butyl [curve (b)] analog, the nearest neighbor distance shrinks from ∼11 to
7 Å. R3PO groups can also be replaced with capping species that are “good leav-
ing groups” such as pyridine to produce a distinct family of entirely inorganic NC
solids. Evaporation of pyridine-capped NC dispersions produces optically clear
thin films in which the pyridine is weakly bound to the NCs. Gently heating the
Annu. Rev. Mater. Sci. 2000.30:545-610. Downloaded from www.annualreviews.org
Access provided by University of Pennsylvania on 04/13/16. For personal use only.

Figure 11 (A) SAXS patterns for close-packed glassy solids of 32 Å NCs with (a) TOPO
caps and (b) TBPO caps. Inset (B) PDFs gives inter-particle separations for (a) of 11 Å
and (b) or 7 Å. (C) SAXS patterns for glassy solids of 34 Å CdSe NCs capped with (c)
pyridine, (d) pyrazine, and (e) bare semiconductor surfaces. Inset (D) PDFs give inter-
particle spacings of (c) 7 Å, (d) 5 Å, and (e) <2 Å.
pyridine-capped NC films to ∼ 150–175◦ C under vacuum quantitatively removes
the pyridine from the thin films, leaving an entirely inorganic superstructure. As
the annealing temperature is raised further, sintering and grain growth occur, ul-
timately producing polycrystalline CdSe thin films (see Exchange Interactions).
Figure 11C, D [curve (c)] shows the scattering from a thin film of pyridine-capped
NCs with a characteristic ∼7 Å inter-particle spacing. Scattering [curve (e)] from
the same film heated under vacuum to remove the pyridine shows that the super-
structure has collapsed, leaving a nanoporous solid. Continued heating leads to
full densification and formation of an entirely inorganic solid as seen in TEM and
by the loss of the SAXS signal.
Annu. Rev. Mater. Sci. 2000.30:545-610. Downloaded from www.annualreviews.org

Cross-linking of the particles in the solid state has the potential to produce mate-
Access provided by University of Pennsylvania on 04/13/16. For personal use only.

rials that are more thermally and mechanically robust. Pyrazine (1,4 azobenzene)
is a bidentate ligand that can bridge neighboring NCs. The SAXS pattern, Figure
11 [curve (d)], indicates that the NCs are separated by 5 Å, consistent with the di-
mensions of single-bridging pyrazine molecules, and is too small to accommodate
independent pyrazine layers on each NC.

Transition from Local to Long-Range Order


Tailoring the composition of the dispersing medium to provide a slow destabi-
lization of the NC dispersion as the solvent evaporates allows the production of
three-dimensional NC superlattices coherent over hundreds of microns (10, 11).
Semiconductor NCs with organic stabilizers having long-chain alkyl tail groups
can be induced to order by evaporating a NC dispersion composed of a low-boiling
alkane and a high-boiling alcohol. As the dispersion is concentrated, the relative
concentration of the alcohol rises, slowly reducing the steric barrier to aggregation
and causing slow separation of the NCs from the dispersed state. Dipolar coupling
between polarizable metal NCs provides an attractive interaction that allows metal
NC superlattices to assemble from single component solvents (57, 114, 115). If
the rate of the transition is carefully controlled, the sticking coefficient between the
NCs remains low and the arrival time of NCs is such that the NCs have sufficient
time to find equilibrium superlattice sites on the growing structure (116, 117). Fig-
ure 12A, B shows schematics of the deposition of glassy and ordered solids, respec-
tively. TEM images (Figure 12C, D) show glassy and ordered solids prepared by
tailoring the solvent composition, as described above, for the same 80 Å CdSe NC
sample.
In the arrival-limited regime, NCs have enough time to diffuse at the growing
surface to form ordered solids. If the arrival rate becomes too large, NCs will
land on top of each other and form an amorphous solid. Cross-sectional TEM
(Figure 13) shows the effects of an increase in NC flux adding to the surface.
Initially, the solvent evaporation rate is slow and the 100 Å NCs deposit as

an ordered NC solid. As the flux of NCs is increased by increasing the evap-


oration rate (i.e. reducing the pressure above the dispersion), the arrival rate
Annu. Rev. Mater. Sci. 2000.30:545-610. Downloaded from www.annualreviews.org
Access provided by University of Pennsylvania on 04/13/16. For personal use only.

Figure 12 Cartoon of the general deposition conditions used in the formation of three-
dimensional, close-packed (A) NC glasses and (B) NC superlattices. TEM images show
that by tailoring the solvent composition, NC samples of 80 Å CdSe NCs, as shown, may
be manipulated into (C ) glassy and (D) ordered NC solids.

exceeds the surface mobility of NCs and the formation of an amorphous solid
occurs.
In an elegant series of experiments, Korgel & Fitzmaurice used real-time SAXS
to reveal the details of the order-disorder transition in NC superlattices (68, 69).
Figure 14 summarizes the real-time SAXS results for the growth of thiol-capped
Ag NC superlattices during evaporation from solution (118). A sharp peak in
the SAXS patterns evolves from a well-defined inter-planar spacing as the NCs
deposit, thus forming a close-packed superlattice. In a second experiment, the
SAXS spectrum of a well-formed Ag NC superlattice is followed as a function of
heating (Figure 15) (119). The disappearance of the superlattice reflections with
heating is consistent with growing disorder as the colloidal assembly melts.
Annu. Rev. Mater. Sci. 2000.30:545-610. Downloaded from www.annualreviews.org
Access provided by University of Pennsylvania on 04/13/16. For personal use only.

Figure 13 TEM cross-section through a h111i-oriented island of 100 Å CdSe NCs. The
sample was cleaved along the h111i axis. The top of the image shows a glassy region
formed at high growth rates. The bottom shows the long-range order possible at slower
growth rates. Striations seen of the (112)SL fcc planes.

Preparation of NC Superlattices (Colloidal Crystals)


When colloidal or NC dispersions become unstable by introducing a nonsol- vent,
the colloidal particles or NCs aggregate and precipitate from solution. The
structure of the aggregates depends on the rate of destabilization and the sticking
coefficient between particles. Very rapid destabilization forms low-density frac-
tal aggregates as particles quickly add and stick to one another (116, 117). At
slower rates of destabilization, the particles form close-packed but amorphous ag-
gregates. Slowly destabilizing the dispersion, for example, by evaporation from a
mixture of solvents and nonsolvents (as described above), results in ordered super-
lattices, also known as colloidal crystals, that homogeneously nucleate in solution.
Mild destabilization provides a weakly attractive potential and sufficient time for
Annu. Rev. Mater. Sci. 2000.30:545-610. Downloaded from www.annualreviews.org
Access provided by University of Pennsylvania on 04/13/16. For personal use only.

Figure 14 Time-resolved SAXS studies of a Ag superlattice as it self-assembles during


evaporation from toluene. The sharp peaks that develop are from the inter-planar spacings
of the deposited Ag superlattice (68).

colloidal particles or NCs to find equilibrium superlattice sites. Ordered NC su-


perlattices have been prepared from semiconductor NCs such as CdSe (10, 11, 45),
InP, AgS (30), and TiO2 (121); from metal NCs such as Au (114, 122), Ag
(57, 68, 69, 123), and Co (31, 124); and from metal-oxide NCs such as Fe2O3
(101, 102), CoO (125), and BaCrO4 (126).
HRSEM captures the three-dimensional morphology of NC structures, which
reveals the stages of superlattice growth (10, 11, 45). Figure 16a shows the ini-
tial stages of crystal deposition as ∼750 62 Å CdSe NCs assemble, forming an

ordered island. The center of the island shows the second and third layer of NCs
adding in three dimensions. Figure 16b shows a larger island of 64 Å NCs as

the NCs add to form a more extensive, ordered island. Colloidal crystals having
regular geometries nucleate homogeneously in solution before precipitating onto
the substrate surface. The most common shapes of the colloidal crystals are in-
complete h111iSL-oriented pyramids (Figure 16c) (SL is used to identify directions
and planes of the superlattice). Within a crop of crystals, the size and shape of
the crystals are very similar. The crystals show terraces and ledges closing off to
form low-index faces. Within a crop of colloidal crystals, <1% square-pyramids
Figure 15 Time-re-
solved SAXS studies of
melting a Ag NC super-
lattice as a function of
temperature. The disap-
pearance of the sharp su-
perlattice reflection upon
heating indicates an order-
to-disorder transition as
the Ag NC superlattice is
amorphized (69).
Annu. Rev. Mater. Sci. 2000.30:545-610. Downloaded from www.annualreviews.org
Access provided by University of Pennsylvania on 04/13/16. For personal use only.

are also seen to nucleate (Figure 16d ). Slower growth rates produce occasionally
larger complete, colloidal crystals with regular geometries. Figure 16e shows a
∼1.7 µm pyramidal-shaped colloidal crystal of 48 Å CdSe NCs. The pyramidal
morphology is characteristic of a h111iSL-oriented face-centered cubic (fcc) crystal
structure. The ledges and terraces have closed off to form vicinal (100)SL facets.
The inset shows a square colloidal crystal characteristic of a h100iSL-oriented fcc
structure.
Optical micrographs show a crop of three-dimensional colloidal crystals of
57 Å CdSe NCs (Figure 17, see color insert) (127). Each triangle is ∼50 µm

on a side. This is an early example of patterning (see Templating and Pattern-


ing of NC Assemblies) (11). Chemically and mechanically patterning surfaces
leads to the spatial organization of colloidal crystals on the substrate. The tri-
angles deposited in lines extend radially outward from the center on the bottom
of a glass vial. Mechanical stresses created in the walls of the glass vial during
Annu. Rev. Mater. Sci. 2000.30:545-610. Downloaded from www.annualreviews.org
Access provided by University of Pennsylvania on 04/13/16. For personal use only.

Figure 16 HRSEM images captures the morphology of self-assembled, close-packed


islands and three-dimensional colloidal crystals of CdSe NCs. (a) The initial stages of
growth for an island of ∼750 62 Å CdSe NCs. (b) Three-dimensional growth forms more
extensive islands of ordered 64 Å CdSe NCs. (c) One in a crop of similar size, incomplete
h111iSL-oriented colloidal crystals of 57 Å CdSe NCs. (d ) At <1% in a crop of colloidal
crystals in the shape of a square pyramid may be found. (e) Slow growth rates form
complete, regular colloidal crystals. Colloidal crystals of 48 Å NCs show the characteristic
pyramidal shape of a h111iSL-oriented fcc structure. Ledges and terraces have closed off,
forming vicinal (100)SL facets. The inset shows a h100iSL-oriented colloidal crystal from
the same sample preparation (45).
Annu. Rev. Mater. Sci. 2000.30:545-610. Downloaded from www.annualreviews.org
Access provided by University of Pennsylvania on 04/13/16. For personal use only.

Figure 17 Optical micrograph of three-dimensional colloidal crystals of 57 Å CdSe QDs.


The triangles form spokes that extend radially outward from the center on the bottom of a
glass vial. The red color of the triangles is characteristic of the size-dependent absorption
for the 57 Å CdSe NC building blocks (127).
575 MURRAY ¥ KAGAN MONODISPERSE
¥ BAWENDI NANOCRYSTAL ASSEMBLIES 575

pulling may have produced surface topography that influenced the deposition of
the colloidal crystals. Similarly, NC superlattices have been seen to preferen-
tially nucleate in regions where the surface has been modified by either chemistry
or abrasion (11). The red color of the triangles is true, arising from the optical
spectrum for the 57 Å CdSe NC building blocks.
HRTEM shows that these colloidal crystals are single NC supercrystals (10).
An image of one in a crop of micron size colloidal crystals is shown in Figure 18A.
The star-shaped pattern in the center of the structure is an electron-channeling
pattern of the h111iSL projection. This pattern clearly reveals the structure to be a
Annu. Rev. Mater. Sci. 2000.30:545-610. Downloaded from www.annualreviews.org
Access provided by University of Pennsylvania on 04/13/16. For personal use only.

single colloidal crystal. An expanded view of one edge of the supercrystal (Figure
16B) shows the 60 Å CdSe NC building blocks of these three-dimensional colloidal
crystals.
Deposition of NC dispersions at an interface maximizes heterogeneous nu-
cleation and allows the production of ordered and oriented epitaxial thin films.
Approximately 80% of the ordered domains are organized with their close-packed
(111)SL planes parallel to the substrate. HRSEM (Figure 19A) shows an ordered
and oriented epitaxial thin film of close-packed 64 Å CdSe NCs. The NCs in the
film show terraces, ledges, and kinks analogous to the Burton, Cabrera, and Frank
model for adatoms adding to a growing surface (128, 129). A cartoon (Figure 19B)
representing the classic terrace-ledge-kink model for monatomic crystal growth is
shown below. Preferential addition of NCs to the steps and ledges produces low-
index planes on extended faces. A second example of NC superlattices is seen in
Figure 19C for 60 Å InP NCs (120). Preparation of monodisperse NC samples

makes the deposition of ordered NC superlattices from semiconductor and metal


NC building blocks possible.
Although ∼ 80% of the domains in epitaxial thin films are h111iSL-oriented with
respect to the substrate surface, extensive surveys of plan-view and cross-sectional
TEM samples help to build a model of the three-dimensional NC superlattices
(10, 45). Two-dimensional projections along the h111iSL, h100iSL, and h110iSL of
a representative fcc superlattice of 48 Å CdSe NCs are shown in Figure 20A, D, G.
Higher magnification images of the superlattice for each projection (Figure 20B,
E, H), resolve the internal lattice of each NC in the fcc superlattice. Small-angle
electron scattering patterns for 2-µm regions of the superlattice are also shown for
each projection (Figure 20C, F, I), confirming the assignment of the superlattice
orientations.
SAXS is used to develop a more precise description of the average structure in
these epitaxial superlattices. Figure 21A shows the sharp superlattice reflections
for films assembled from a size series of NC samples. Patterns indicate a strong,
preferential alignment of the close-packed h111iSL parallel to the substrates, as seen
in SEM and TEM. Control over the spacing between NCs is achieved through cap-
exchange. Figure 21B shows the tunability in inter-particle separation for 34 Å
CdSe NCs with R3PO/R3P caps. For R = hexadecyl, octyl, and butyl moieties,
the inter-particle spacing is tuned from 17 to 11 to 7 Å, respectively.
576 MURRAY ¥ KAGAN MONODISPERSE
¥ BAWENDI NANOCRYSTAL ASSEMBLIES 576
Annu. Rev. Mater. Sci. 2000.30:545-610. Downloaded from www.annualreviews.org
Access provided by University of Pennsylvania on 04/13/16. For personal use only.

Figure 18 (A) TEM image of a faceted colloidal crystal prepared from 60 Å CdSe NCs.
A star-shaped electron channeling pattern along the h111iSL dramatically demonstrates
ordering in the single domain colloidal crystal. (B) TEM image of one facet of the colloidal
crystal showing the 60 Å CdSe NC building blocks.
577 MURRAY ¥ KAGAN MONODISPERSE
¥ BAWENDI NANOCRYSTAL ASSEMBLIES 577
Annu. Rev. Mater. Sci. 2000.30:545-610. Downloaded from www.annualreviews.org
Access provided by University of Pennsylvania on 04/13/16. For personal use only.

Figure 19 (A) HRSEM of a three-dimensional superlattice of 64 Å CdSe NCs grown

epitaxially on an amorphous carbon substrate. Ordered rows of NCs form terraces, ledges,
and kinks. (B) Superlattice growth is analogous to the terrace, ledge, kink model that
describes monotonic crystal growth. (C ) TEM image shows an ordered assembly of 60 Å
InP NCs (120).

Preferred Orientation of NCs in Superlattices


The optical and electronic properties of NC assemblies depend intimately on the
orientation of the constituent NCs and their inherent anisotropy. Several experi-
mental methods have been realized to produce oriented nanometer size semicon-
ductors. Lattice mismatch between a substrate and a vacuum-deposited overlayer
by, e.g. molecular beam epitaxy or chemical vapor deposition, gives rise to strain
that induces growth of nanometer size islands. Oriented islands self-organize
into superlattices growing under the influence of the lattice strain (130). An-
other approach is to implant constituent semiconductor ions into crystalline solid
substrates followed by annealing, which results in the nucleation and growth of
crystallographically oriented semiconductor NC inclusions (131).
578 MURRAY ¥ KAGAN MONODISPERSE
¥ BAWENDI NANOCRYSTAL ASSEMBLIES 578
Annu. Rev. Mater. Sci. 2000.30:545-610. Downloaded from www.annualreviews.org
Access provided by University of Pennsylvania on 04/13/16. For personal use only.

Figure 20 (A) TEM image of a h100iSL projection through a three-dimensional super-


lattice of 48 Å CdSe NCs. (B) At high magnification, the internal lattice structure of the
NC building blocks is resolved. (C ) Small-angle electron diffraction pattern demonstrates
the lateral perfection of the superlattice domain. (D) TEM image showing the h101iSL
projection for a superlattice of 48 Å CdSe NCs. (E ) High magnification shows lattice

imaging of the individual NCs. (F ) Small-angle electron diffraction demonstrates the


perfection of another characteristic fcc orientation. (G) TEM image of an fcc superlattice
of 64 Å CdSe NCs viewed along the h111iSL axis. Inset (H ) higher magnification image
showing the individual NCs in the superlattice and (I ) a small-angle electron diffraction
pattern showing ordering in the h111iSL orientation over the probed 2 µm area.
579 MURRAY ¥ KAGAN MONODISPERSE
¥ BAWENDI NANOCRYSTAL ASSEMBLIES 579
Annu. Rev. Mater. Sci. 2000.30:545-610. Downloaded from www.annualreviews.org
Access provided by University of Pennsylvania on 04/13/16. For personal use only.

Figure 21 (A) SAXS patterns for epitaxial superlattices prepared from CdSe NCs ranging
from 35 to 64 Å in diameter. The fcc superlattice relections are indicated for the 64 Å CdSe
NC superlattice. (B) SAXS patterns of ordered epitaxial thin films prepared from 34 Å
CdSe NCs derivatized with (e) tricetylphosphate, giving a ∼17 Å inter-particle spacing;
with (f ) the native trioctylphosphine chalcogenide, giving a ∼11 Å spacing; and with
(g) tributylphosphine oxide, giving a ∼7 Å spacing (45).

During the growth of epitaxial NC superlattices the individual NCs align them-
selves to maximize their attractive interaction (van der Waals and dipolar) (132)
with the substrate and neighboring NCs. The TEM image (Figure 22A) shows
even more precise alignment of the NCs. This image has a h111iSL projection of a
superlattice of 64 Å NCs grown epitaxially on a TEM grid. Clearly visible columns
of atoms produce a fine cross-hatched pattern throughout the image. Columns of
Cd and Se atoms in a given NC are coherent with those of their neighbors over
the area of observation. Alignment of the h002i axis of NCs parallel to the sub-
strate enhances the intensity of the h002i reflection seen in Figure 22B. Hexagonal
modulation of the ring intensity is due to the coherent alignment of the individual
580 MURRAY ¥ KAGAN MONODISPERSE
¥ BAWENDI NANOCRYSTAL ASSEMBLIES 580
Annu. Rev. Mater. Sci. 2000.30:545-610. Downloaded from www.annualreviews.org
Access provided by University of Pennsylvania on 04/13/16. For personal use only.

Figure 22 (A) High resolution TEM image shows a section of a h111iSL-oriented su-
perlattice of 62 Å CdSe NCs. The periodic dot pattern running through the image arises
from the coherent imaging of columns of Cd and Se atoms making up the NCs in the
superlattice. (B) Wide-angle electron diffraction pattern from a ∼2 µm area. Strong mod-
ulation of the diffraction pattern results from the preferred alignment of the individual NC
axes within the superlattice. At lower flux, the small-angle electron diffraction pattern
can be seen emerging from the central beam. (C ) Wide-angle electron diffraction pattern
from a NC glass, having an isotropic orientation of NCs, is prepared from the same NC
sample.
581 MURRAY ¥ KAGAN MONODISPERSE
¥ BAWENDI NANOCRYSTAL ASSEMBLIES 581
Annu. Rev. Mater. Sci. 2000.30:545-610. Downloaded from www.annualreviews.org
Access provided by University of Pennsylvania on 04/13/16. For personal use only.

Figure 23 SAXS and WAXS for (a) an oriented epitaxial and (b) a glassy thin film prepared
from 64 Å CdSe NCs. The ordering and orientation of the NCs in the epitaxial thin film
are evidenced by the sharp reflections in SAXS and the dramatic enhancement in the (110)
reflection in WAXS (45).

NC’s wurtzite lattice with the fcc superlattice. A blowup of the central beam
(inset, Figure 22B) at lower flux shows the small-angle electron diffraction pattern
emerging. For comparison, Figure 22C shows the electron diffraction pattern of a
glassy film prepared from the same NC sample that displays the uniform wurtzite
ring pattern of an isotropic distribution of NCs. Similar oriented NC superlattices
have been prepared from metal (133) and metal-oxide (125) NCs.
The effects of NC orientation within superlattices are also seen in X-ray diffrac-
tion (10, 45). SAXS scattering patterns (left), shown in Figure 23, compare and
contrast diffraction from (a) an ordered epitaxial thin film and (b) a glassy thin
film prepared from the same sample of 64 Å CdSe NCs. The SAXS pattern from
epitaxial thin films is dominated by the strong reflections off the ordered superlat-
tice planes, while that for the glassy thin film is dominated by the oscillatory form
factor of its NC building blocks. The WAXS portion of the diffraction (right) for
the glassy thin film [curve (b)] is consistent with a randomly oriented ensemble
of NCs. The WAXS pattern (right) for the epitaxial thin film [curve (a)] shows a
striking enhancement of the wurtzite (110) reflection. The intensity of the (110)
reflection indicates a greater that 90% alignment of the wurtzite c-axis of the
individual NCs in the plane of the substrate.
582 MURRAY ¥ KAGAN MONODISPERSE
¥ BAWENDI NANOCRYSTAL ASSEMBLIES 582

Binary NC Assemblies
Binary assemblies of submicron colloidal particles have been investigated theo-
retically and experimentally. Bimodal assemblies of larger silica and polymeric
spheres have been observed to form intermetallic phases such as AB2 and AB13
(134–136). Recently, the first clear examples of binary assembly in the <100 Å
diameter size region was observed by Schiffrin and co-workers for thiol stabilized
Au NCs (115). These binary assemblies are also similar to intermetallic phases
found in alloys between metals having different atomic radii. For Au NCs with
Annu. Rev. Mater. Sci. 2000.30:545-610. Downloaded from www.annualreviews.org
Access provided by University of Pennsylvania on 04/13/16. For personal use only.

radius ratios of ∼0.58, the intermetallic AB2 phase forms. In the example, the
smaller 45 Å Au NCs fit in the interstices of the hexagonally arranged larger 78 Å
Au NCs. Mixing two different size NC samples with an appropriate ratio in size
and packing fraction to form an intermetallic phase produces NC assemblies with
higher packing densities than single component fcc or hcp assemblies.
For mixtures of NCs with sizes outside the range of radius ratios for an in-
termetallic phase, the NC assemblies either phase-separate, forming regions of
smaller and larger NCs, or randomly mix, forming an alloy of small and large
NCs. The arrangement of the NCs in the solids also depends on the degree to
which the NCs order during deposition (11, 67). Figure 24a shows an ordered NC
solid prepared from a mixture of 82% 37.5 Å and 18% 57 Å CdSe NCs in a mixture
of low-boiling alkane (solvent) and high-boiling alcohol (nonsolvent). The small
and large NCs have phase-separated into ordered regions. Figure 24b shows that
in a glassy NC solid, prepared from a similar mixture of NCs, the small and large
NCs remain intermixed. The repulsive interaction between all NCs is maintained
as the dispersion evaporates until it solidifies. Binary assemblies of either two
different sizes or two different compositions of NCs is an interesting future di-
rection for both understanding how to control their assembly and how to prepare
materials that are periodically electronically modulated on the nanometer length
scale.

Templating and Patterning of NC Assemblies


The ability to manipulate NCs and NC assemblies into robust integrated struc-
tures will be essential to their application in any future technology. Devices that
exploit the novel electronic properties of NCs and their assemblies will require
fabrication of NC and NC assemblies with precise spatial control. Conventional
optical lithography permits templating of NC assemblies. Surfaces derivatized
with monolayers that bind strongly to the NCs have been used as a route to selec-
tively deposit NC assemblies. Vossmeyer and co-workers used optical lithography
to photo-deprotect regions of a monolayer containing nitroveratryloxycarbonyl-
glycine (NVOC-Gly) end groups (137, 138). When exposed to light the NVOC
functionality is cleaved, generating a glycine-rich surface whose terminal amines
are available for binding to the NCs. Deposition of an Au NC dispersion results
in preferential binding of the NCs to the exposed regions.
583 MURRAY ¥ KAGAN MONODISPERSE
¥ BAWENDI NANOCRYSTAL ASSEMBLIES 583
Annu. Rev. Mater. Sci. 2000.30:545-610. Downloaded from www.annualreviews.org
Access provided by University of Pennsylvania on 04/13/16. For personal use only.

Figure 24 Tailoring the composition of the solvents to deposit ordered or glassy binary
assemblies of two different sizes of NCs. (a) HRSEM showing a binary assembly of 82%
small, 37.5 Å and 18% large, 57 Å CdSe NCs. The mixture of low-boiling alkane and
high-boiling alcohol deposits ordered, phase-separated regions of the small and the large
NCs. (b) Using good solvents that maintain a repulsive inter-particle interaction during
solvent evaporation, HRSEM shows, for the binary mixture of 82% 37.5 Å and 18% 62 Å
CdSe NCs, that the two sizes of NCs remain well-intermixed forming a close-packed glassy
solid (67).
584 MURRAY ¥ KAGAN MONODISPERSE
¥ BAWENDI NANOCRYSTAL ASSEMBLIES 584
Annu. Rev. Mater. Sci. 2000.30:545-610. Downloaded from www.annualreviews.org
Access provided by University of Pennsylvania on 04/13/16. For personal use only.

Figure 25 Low (a) and high (b) magnification SEM images show the internal structure
of a three-dimensionally ordered macroporous solid. A sintered, glassy CdSe NC solid
provides a high refractive index scaffolding in contrast to that of the pores. The Fourier
transform of the image shows that the pores have long-range order in the h111i direction of
the solid (139).

Recently, Norris and co-workers used CdSe NCs as building blocks to produce
ordered three-dimensional porous solids of high- and low-refractive indexes as
potential photonic bandgap materials (139). The CdSe NCs are patterned in three-
dimensions by immersing a prefabricated synthetic silica opal in a NC-dispersion
tailored to produce a glassy solid upon evaporation. The NCs are sintered, which
forms an inorganic solid, as described above. The silica spheres are selectively
etched from the solid, which produces a solid whose refractive index is modulated
in three dimensions, between nair = 1 and nCdSe ∼ 2.7, on the length scale of light.
Figure 25 shows low and high magnification SEM images of the remaining CdSe
585 MURRAY ¥ KAGAN MONODISPERSE
¥ BAWENDI NANOCRYSTAL ASSEMBLIES 585

scaffolding. The three-dimensional ordering of pores is seen by the sharp spots in


the Fourier transform of the image (inset Figure 25a).
Efforts are underway to combine the specificity of DNA and the assembly
of NCs to develop new NC architectures (140–142). Examples now exist of
derivatizing the surface of semiconductor and metal NCs with segments of single-
stranded DNA. By introducing the derivatized NCs in a dispersion containing
complementary strands of DNA, the NCs attach to the DNA strands. Adjusting
the repeat sequence of the complementary DNA strand offers a route to control
the position at which the derivatized NCs attach such that NCs are attached either
Annu. Rev. Mater. Sci. 2000.30:545-610. Downloaded from www.annualreviews.org
Access provided by University of Pennsylvania on 04/13/16. For personal use only.

on the same side or opposite sides of the DNA strand.

OPTICAL AND ELECTRONIC PROPERTIES


OF NC ASSEMBLIES
Introduction
Building close-packed solids of semiconductor or metal NCs presents opportu-
nities to explore cooperative physical phenomena that develop as proximal NCs
interact and to engineer the electronic, optical, optoelectronic, and magnetic prop-
erties of materials on a nanometer scale. Understanding the interactions between
NCs is the first step in designing the properties of NC solid-state materials and
devices. Tailoring the size and composition of the NCs and the length and elec-
tronic structure of the matrix may be used to tune inter-particle couplings and the
properties of NC assemblies.
Semiconductor NCs are also known as quantum dots (QDs). QD or NC solids
represent zero-dimensional analogs of quantum well (QW) superlattices. Under-
standing and electronically engineering the couplings between QWs in heterostruc-
tures has and continues to be of interest for both fundamental physics and device
applications. Similarly, solids of metal NCs are zero-dimensional analogs of metal-
lic multilayers. Close-packed NC assemblies provide a testing ground to study
quantum mechanical effects and materials physics as the electronic structure of
these solids may be modulated in three dimensions on a length scale compara-
ble to the mean free path and natural extent of electronic carriers and excitations
(143, 144). These structures may provide insight into the physical interactions that
develop as atoms or NCs approach and that give rise to the electronic and opti-
cal properties of solid-state materials. NC solids are also analogous to molecular
solids because many of the properties of individual NCs are comparable to large
molecules and because NCs and molecules are similarly bound by weak van der
Waals forces in the solid state. NC solids combine both the unique atomic- and
molecular-like properties of individual NCs and the collective solid-like properties
of coupled NCs. This combination of materials properties makes close-packed NC
solids interesting media for potentially novel electronic, optical, and optoelectronic
device applications.
586 MURRAY ¥ KAGAN MONODISPERSE
¥ BAWENDI NANOCRYSTAL ASSEMBLIES 586

Weak Coupling
One possible interaction between proximal atoms, molecules, or NCs is dipolar
coupling. Electric dipoles can be either permanent dipoles resulting from per-
manent displacement of electric charge, particularly found in highly polarizable
molecules or NCs, or transition dipoles arising from oscillations in charge dis-
tributions as atoms, molecules, or NCs change eigenstate. Interactions between
permanent and transition dipoles have been calculated theoretically, and the case
of transition dipoles has been shown experimentally. Dipole-dipole interactions
Annu. Rev. Mater. Sci. 2000.30:545-610. Downloaded from www.annualreviews.org
Access provided by University of Pennsylvania on 04/13/16. For personal use only.

occur when NCs are placed in proximity and give rise to electronic and optical
properties unique to the solid state.
Interactions between permanent electric dipoles and their formation have been
theorized in NC assemblies because individual NCs are highly polarizable (136).
Organizing NCs into solids is expected to induce permanent electric dipoles as
neighboring NCs polarize one another (145, 146). This spontaneous polarization,
unique to the solid state, has been modeled and is expected to lead to the formation
of ferroelectric and antiferroelectric phase transitions in NC arrays (147). Forma-
tion of either ferroelectric or antiferroelectric phases is calculated and depends on
the geometry of the NC array. Interactions between permanent electric dipoles
are expected to manifest themselves as high frequency, collective modes that ap-
pear at far-infrared frequencies in optical absorption (148, 149). Applications of
dipolar coupling have also been proposed that aim at using NC solids as media
for complex computations. In this scheme, dipole-dipole interactions act as wire-
less substitutes for electrical interconnects in information transfer, mimicking the
design of cellular automata (150, 151).
Coupling between transition dipoles is known to occur when an excited atom,
molecule, or NC is placed in the near-field of a ground state atom, molecule, or
NC. At inter-atomic, inter-molecular, or inter-particle separations between 5 and
100 Å, dipole-dipole interactions lead to electronic energy transfer, also known
as long-range resonance transfer (LRRT) (67, 152–157). LRRT is a radiation-
less energy transfer process that arises as transition dipoles of resonant transitions
in an excited NC (or atom or molecule) and a ground state NC couple through
their generated electromagnetic fields. This coupling causes the excitation in the
excited NC (known as the donor of the excitation) to be transferred to the ground
state NC (known as the acceptor of the excitation), returning the excited NC to its
ground state and promoting the ground state NC to a higher excited state. LRRT
is a well-known phenomenon occurring between molecules in organic solids and
between impurity centers in inorganic solids (153–155).
LRRT is characterized by a very weak coupling of transition dipoles so typical
rates of excitation transfer (∼ 10−8 s) are much smaller than the rates of absorption
(∼10−14 s) and vibrational relaxation (∼ 10−12 s) processes in individual NCs.
Figure 26 shows optical absorption and PL spectra for a representative series of
close-packed NC solids prepared from CdSe NC samples ranging in size from
30 to 62 Å in diameter (11, 67). The TOPO/TOP ligands coordinating the NCs’
587 MURRAY ¥ KAGAN MONODISPERSE
¥ BAWENDI NANOCRYSTAL ASSEMBLIES 587
Annu. Rev. Mater. Sci. 2000.30:545-610. Downloaded from www.annualreviews.org
Access provided by University of Pennsylvania on 04/13/16. For personal use only.

Figure 26 Optical absorption and photoluminescence spectra at 10 K of optically thin


and clear, close-packed NC solids prepared from CdSe NC samples (A) 30.3, (B) 39.4,
(C) 48.0, and (D) 62.1 Å in diameter (67).
588 MURRAY ¥ KAGAN MONODISPERSE
¥ BAWENDI NANOCRYSTAL ASSEMBLIES 588

surfaces maintain ∼11 Å inter-particle separation. The discrete optical absorption


resonances and sharp band-edge emissions are characteristic of the size-dependent,
quantized electronic transitions for these same CdSe NC samples dispersed in
solution. Absorption of light generates electronic excitations that are completely
localized within the individual NCs in the solids. Although the absorption spectra
remain unchanged, as discussed below, the emission spectra for close-packed and
dispersed NCs differ and can be understood by accounting for energy transfer
between NCs in the solids.
Annu. Rev. Mater. Sci. 2000.30:545-610. Downloaded from www.annualreviews.org
Access provided by University of Pennsylvania on 04/13/16. For personal use only.

Mixed NC Solid In solids containing two dissimilar molecules or two sizes of


NCs, one the donor and the other the acceptor, LRRT is measured spectroscopically
by the quenching of the PL QY or the decrease in the PL lifetime of the donor
or by the accompanied enhancement of the PL QY or increase in the PL lifetime
of the acceptor (11, 67, 152–157). For LRRT to occur and be observed, two
conditions must be satisfied: The acceptor must have both a transition in resonance
with the energy of donor emission and a lower energy state in which to trap the
transferred excitation (153–157). Once the transferred excitation is trapped (at
rates typical of vibrational relaxation) in the lower energy state of the acceptor,
the excitation cannot be transferred back because the donor is transparent to the
lower energy excitation. Examples of LRRT in organic systems containing two
dissimilar molecules include combinations of oligocenes. For example, tetracene,
which has a longer conjugation length, acts as an acceptor in solids of its shorter
conjugated analog, anthracene, which acts as the donor (153).
By taking advantage of the size-dependent electronic spectrum of NCs, a mixed
NC solid containing 82% 38.5 Å NCs (small NCs) and 18% 62 Å NCs (large

NCs) provides a test system to demonstrate dipole-dipole interactions between


proximal NCs (11, 67, 152). The mixed NC solid is deposited as a well-intermixed,
glassy solid, as is shown by HRSEM (Figure 24b). The mixed system satisfies
the conditions required to observe energy transfer. As depicted in the cartoon
(Figure 27), the small, 38.5 Å NCs are the donors and the large, 62 Å NCs are

the acceptors. The large NC has both a transition (|g> → |A1>) resonant with
the emission of the small NC (|D> → |g>) and a lower energy state (|A2>) in
which to trap the transferred excitation. The small NC is transparent to this lower
energy.
Optical absorption spectra at RT and 10 K are shown in Figure 28a, b for the
mixed NC solid. The absorption spectrum for the mixed NC solid is a linear
combination of the absorption spectra for the small and large NCs. Subtracting
the spectrum for the large NCs (dotted lines) produces that for the small NCs.
RT and 10 K PL spectra for the NCs close-packed in the solid and dispersed
in solution are shown as solid lines in Figure 28c–f for excitation at 2.762 eV,
(arrow 1 in Figure 28a, b). A dramatic increase in the ratio of PL intensity for
the large to small NCs is observed in the solid state. The relative PL QY for a
pure solid of the same small, 38.5 Å CdSe NCs, with no large NCs, is plotted
589 MURRAY ¥ KAGAN MONODISPERSE
¥ BAWENDI NANOCRYSTAL ASSEMBLIES 589
Annu. Rev. Mater. Sci. 2000.30:545-610. Downloaded from www.annualreviews.org
Access provided by University of Pennsylvania on 04/13/16. For personal use only.

Figure 27 Cartoon for electronic energy transfer between two different size NCs in a
close-packed NC solid. The small NCs, having larger effective bandgap, are the donors
of excitations; the larger NCs, having smaller effective bandgap, are the acceptors. The
large NCs have both an absorption (|g>→|A>1) in resonance with the small NC emission
(|D>→|g>) and a lower energy transition (|A>2→|g>) in which to trap the transferred
excitation. Energy transfer provides a competing pathway, with rate kDA, to radiative and
nonradiative relaxation in the small NCs, with rates kDR and kDNR · kAR and kANR the radiative
and nonradiative rates of relaxation in the large NCs (67).

(dotted line) under label A. Excitation of the mixed NC solid to the red of the
small NCs absorption (arrow 2) (2.143 eV) excites only the large NCs in the
mixed solid. The relative PL QY for the large NCs is shown (dotted line) under
label B. The PL spectra, both at RT and 10 K, for the mixed NC solid are not the
same weighted combination of small and large NC PL spectra. The PL QY for the
small NCs is quenched and the PL QY for large NCs, upon exciting both small and
large NCs, is enhanced in the mixed solid. These observations are signatures of
electronic energy transfer from the small to the large NCs in the close-packed NC
solid.
Förster calculated the efficiency of LRRT by relating the dipole-dipole inter-
action to spectroscopically accessible measurements of donor emission and ac-
ceptor absorption (153–157). The efficiency is expressed as a distance, known as
the critical distance (Ro), compared with the physical distance, RDA, separating
the donor and acceptor in the solid. Ro is the distance at which the rate of en-
ergy transfer (kDA ) is equivalent to the rate of donor de-excitation by competing
radiative (kDR ) and non-radiative (kDNR ) relaxation processes (Figure 27). For
a random orientation of transition dipoles (153–157), as for NCs in glassy
Annu. Rev. Mater. Sci. 2000.30:545-610. Downloaded from www.annualreviews.org
Access provided by University of Pennsylvania on 04/13/16. For personal use only.

Figure 28 (a) RT and (b) 10 K optical absorption spectra for a glassy NC solid of 82%
37.5 Å (small) and 18% 62 Å (large) CdSe NCs (solid line). The absorption and PL features
of the small and large NCs in the mixed NC solid are spectrally distinct. Subtracting the
absorption spectrum for the large NCs (dotted line) produces the absorption spectrum for
the small NCs (dashed line). Solid lines show PL spectra for excitation at 2.762 eV [arrow
1 in (a, b)] of the same mixed NC sample dispersed in solution, at (c) RT and (d ) 10 K, and
close-packed into a glassy solid, at (e) RT and ( f ) 10 K. Quenching of the small NCs PL
QY in the mixed solid compared with that in a pure, small NC solid (dotted line under label
A) and enhancement of the large NCs PL QY in the mixed solid for excitation at 2.762 eV
(arrow 1) compared with excitation at 2.143 eV (arrow 2) are signatures of energy transfer
from the small to the large NCs is the solid state (67).
591 MURRAY ¥ KAGAN MONODISPERSE
¥ BAWENDI NANOCRYSTAL ASSEMBLIES 591

solids (11, 67, 152),


µ Z ∞ ¶1/6
ϕD dν̃
Ro ∝ FD (ν̃ )εA (ν̃ ) 4 , 8.
n4 0 ν̃

where ϕ D is the PL QY of the donor and n is the refractive index of the medium.
The efficiency of energy transfer can be understood from its origin in dipolar
coupling. The integral is the spectral overlap of donor emission [expressed as
the normalized emission spectrum FD (ν)] and acceptor absorption [expressed as
the molar absorption coefficient εA (ν)] that represents the resonance condition
Annu. Rev. Mater. Sci. 2000.30:545-610. Downloaded from www.annualreviews.org
Access provided by University of Pennsylvania on 04/13/16. For personal use only.

required between donor and acceptor transitions. Using Equation 8, Ro = 47 Å


at RT and 67 Å at 10 K. The increase in energy transfer efficiency with decreasing
temperature originates from the increase in NC PL QY at low temperature.
These values of Ro are reproduced by two independent spectroscopic measure-
ments (11, 67, 152). One is from the quenching of the donor PL QY in the mixed
NC solid compared with that for the pure solid of small NCs. The other is from
time-resolved PL measurements. PL decays show the expected decrease in life-
time for the small NCs in the mixed NC solid from that in a pure solid. This is
accompanied by the similarly expected increase in the PL lifetime for the large
NCs upon exciting both the small and large NCs in the mixed NC solid. Compar-
ing Ro to RDA (RDA ∼ 61 Å) reveals that dipole-dipole interactions are significant
only between nearest neighbor NCs in the solids. The probability of electronic
energy transfer (PDA) calculated from Ro and RDA, using (158)
Ro6
PDA = , 9.
6
o+R
is 0.17 at RT and 0.63 at 10 K. R6 DA

Electronic energy transfer in mixed NC solids is not unique to the size or com-
position of the NCs or to the chemistry of the organic capping group. Energy
transfer between other combinations of NC sizes in close-packed solids is also ob-
served, but with an efficiency dependent on the spectral overlap of donor emission
and acceptor absorption and ϕ D and n for the particular NC samples. Micic et al
have reproduced the effects of energy transfer in a similar mixed system prepared
from InP NCs (159). Electronic energy transfer is also observed in mixed CdSe
NC samples having similar ligands, but with shorter tri-butyl chains and much
longer tri-octadecyl chains. These differences in separation between NC surfaces
for the different length capping groups have only a small effect on the probability
of energy transfer. The relevant physical distance, RDA, between donor and accep-
tor is the distance between the centers of neighboring NCs and not their surfaces
(see Higher Multipole Interactions).

Single-Size NC Solids Even in size-selected NC samples with size distributions


<5%, there remains an inhomogeneous distribution in NC emitting energies. The
spectral inhomogeneity arises not only from the residual size distribution in NC
samples but also from environment. This inhomogeneous distribution has been
592 MURRAY ¥ KAGAN MONODISPERSE
¥ BAWENDI NANOCRYSTAL ASSEMBLIES 592

shown in single NC PL measurements as the PL spectra for NC samples can be


rebuilt from the sum of structured single-NC PL spectra (88).
Differences in the PL spectra for close-packed and dispersed NC samples
(Figure 29) can be understood as energy is transferred from NCs with electronic
transitions on the higher energy side to NCs with electronic transitions on the
lower energy side of the distribution (11, 67). As energy is transferred from high
to low energy, the PL spectra for NC solids, compared with that of dispersed NC
samples, are red-shifted and the lineshapes become asymmetric, accentuated in
the red, and narrowed. These changes in PL position and lineshape are reversible
Annu. Rev. Mater. Sci. 2000.30:545-610. Downloaded from www.annualreviews.org
Access provided by University of Pennsylvania on 04/13/16. For personal use only.

upon redispersing the NC solid. This red shift originates from energy transfer
between NCs in the solid and not from a solvent effect. The red shift in PL for
close-packed solids has been observed for solids prepared from CdS (160), CdSe
(11, 67), and InP (159) NC samples. Energy transfer within a spectrally inhomo-
geneous distribution of emitters is also observed in biological systems such as the
photosynthetic bacterium Rhodospirillium rubrum (161, 162).
The 10 K PL spectra for close-packed and dispersed CdSe NC samples are
compared in Figure 29. The magnitude of the red shift for these NC samples
ranges from 15 to 35 meV, varies from sample to sample with no discernible size
dependence, and decreases with increasing temperature. The sample-to-sample
variations and the temperature dependence originate from the energy transfer ef-
ficiency’s (Ro in Equation 8) dependence on PL QY.
To calculate the probability of energy transfer in single size NC solids, using
Equation 8, the initial distributions of emitting energies for the NC samples are
given by the PL spectra for the same NC samples dispersed in solution. In general,
for similar PL QYs, Ro increases as the spectral overlap increases (Stokes shift
between absorption and PL decreases) with increasing NC size. Comparing values
of Ro with the nearest neighbor distances RDA shows that for single size NC solids
energy transfer is also significant only between nearest neighbors. Using Equation
9, PDA = 38%, 11%, 21%, 14% for NC solids A, B, C, and D, respectively. In
general, since RDA increases more rapidly than Ro, the probability of energy transfer
in the solids decreases with increasing NC size.
The spectral changes in PL for close-packed NC solids are reproduced from
their parent solutions by accounting for energy transfer within the inhomogeneous
distribution in emission. The inhomogeneous distribution is a sum of structured,
single NC PL spectra that consist of a narrow band-edge emission and a longitudinal
optical (LO)-phonon progression. The single NC PL is described by
4 µ ¶
1 (se )n [ν −(ν 0−nωLO )]2
X
E(ν, ν 0 ) = √ exp − 2
10.
n=0
2π γn n! 2γn

where ν 0 is the energy of the zero-order LO-phonon line in emission, and Se


is the strength of the exciton-LO-phonon coupling in emission (84, 85). To obtain
593 MURRAY ¥ KAGAN MONODISPERSE
¥ BAWENDI NANOCRYSTAL ASSEMBLIES 593

the characteristics of the initial inhomogeneous distribution for each NC sample,


the PL spectra for samples dispersed in solution are fit, assuming a Gaussian
Annu. Rev. Mater. Sci. 2000.30:545-610. Downloaded from www.annualreviews.org
Access provided by University of Pennsylvania on 04/13/16. For personal use only.

Figure 29 Comparison of 10 K PL spectra shows, for CdSe NCs (A) 30.3, (B) 39.4, (C )
48.0, and (D) 62.1 Å in diameter, that the PL for NCs close-packed into solids are red-shifted
from that of their parent solution. The inhomogeneous distribution of emitting energies for
each NC sample is extracted by fitting the PL spectra for NCs dispersed in solution to a
Gaussian distribution of single NC PL spectra (open squares). The probability of energy
transfer is calculated from the spectral overlap of absorption and PL for NCs solutions. The
changes in emission lineshape (open circles) that occur upon close-packing NCs into solids
(accounting for energy transfer within the inhomogeneous distribution) are shown.
594 MURRAY ¥ KAGAN MONODISPERSE
¥ BAWENDI NANOCRYSTAL ASSEMBLIES 594

distribution D(ν 0 , νo ) of single NC spectra centered at ν o with standard deviation


γ. The summation is taken over the first five LO-phonon replicas that are spaced
by the LO-phonon frequency for CdSe, ωo, and have linewidths γ n· Se and γ n
depend on NC size. Using values for Se and γ n for the sample’s average NC size,
from FLN measurements (84, 85) on dispersed NC samples, the full PL spectra
are fit (Figure 29 open squares) using nonlinear least squares methods to
Z
EPL (ν) = C E(ν, ν 0 )D(ν 0 , νo ) dν 0 . 11.

Single NC PL measurements, while showing similar values for Se, reveal much
narrower emission linewidths (88). Although not shown here, fitting the PL spectra
for dispersed NCs with these narrower linewidths, yields larger standard deviations
for the inhomogeneous distributions. Using these narrower linewidths and larger
standard deviations to calculate the effects of energy transfer, the model described
below returns spectra for the NC solids virtually identical to those shown using
FLN parameters.
Using the probabilities of energy transfer for each NC sample, energy transfer
is simulated between each NC donor and the acceptor, which have lower emission
energy, and are in the donor’s nearest neighbor shell of twelve NCs in these three-
dimensional close-packed solids (11, 67). The number of acceptors, NA, for a NC
donor with emission energy νemi is given by (11, 67)
Z νem · ¸
12 i (ν −νo )2
NA (νemi ) √ exp − dν. 12.
2π γ −∞ 2γ 2

The probability that NCs do not transfer their energy, therefore the initial and final
emission energies are equal (νemf = νemi ), depends on the probability of energy
transfer and the number of nearest neighbor acceptors as
· ¸
1 (νemi −νo )2 NA (νem )
i

D(νemf = νemi ) = √ exp − (1 − PDA ) . 13.


2π γ 2γ 2

The number of acceptors is larger for NCs with emitting energies on the higher en-
ergy side of the inhomogeneous distribution. The probability that NCs do transfer
their energy to a NC with emission energy νemfi is described by
h i h i
¤
12 (νemf −νo )2 (νem −νo ) 2 NA (νemi )

Z ∞ exp − exp − 1−(1−PDA )
2π γ 2 2γ 2 2γ 2

A(νemf ) = dνemi .
νemf NA (ν emi )
14.
595 MURRAY ¥ KAGAN MONODISPERSE
¥ BAWENDI NANOCRYSTAL ASSEMBLIES 595

The emission spectrum from the NC solid is a sum of emissions from NCs that
do not transfer their energy, D(νemf = νemi ), and those that do transfer their en-
ergy, A(νemf ). The calculated spectra accounting for energy transfer within the
inhomogeneous distribution (open circles, Figure 29) for each of the NC samples
reproduces, most obviously, the red shift in emission position measured for NCs
Annu. Rev. Mater. Sci. 2000.30:545-610. Downloaded from www.annualreviews.org
Access provided by University of Pennsylvania on 04/13/16. For personal use only.
596 MURRAY ¥ KAGAN MONODISPERSE
¥ BAWENDI NANOCRYSTAL ASSEMBLIES 596

close-packed into solids. While not shown here, the magnitude of the red shift
and the narrowing of the emission lineshape become more apparent as energy is
transferred within wider inhomogeneous distributions (11, 67).
Applications of close-packed NC solids as nonlinear optical materials have been
proposed (163, 164). The interaction between transition dipoles may be exploited
because it is expected to further augment the already increased optical nonlin-
earity of individual NCs gained by concentrating the bulk oscillator strength in
the discrete electronic transitions of the NC. The dipole-dipole interaction enables
electronic excitations to collect additional oscillator strength from neighboring
Annu. Rev. Mater. Sci. 2000.30:545-610. Downloaded from www.annualreviews.org
Access provided by University of Pennsylvania on 04/13/16. For personal use only.

NCs in the solid through energy transfer.

Higher Multipole Interactions While the dipole-dipole interaction character-


ized by the LRRT model accounts well for the observed changes in optical spectra
due to energy transfer, it is not strictly valid (11, 152). Förster’s LRRT theory is
calculated for point dipoles separated by an effectively infinite distance (149). For
molecules or NCs, the distance between neighbors may be smaller than or com-
parable to the finite volume of the molecule or NC. At distances <20 Å, which is
the case for unit cells near neighboring NC surfaces, higher multipole interactions
between unit cell transition moments may contribute to energy transfer. The case
of finite volumes has been treated theoretically for molecular systems (145). For
NCs, dipolar coupling is expressed as a sum of dipole-dipole interactions between
1
combinations of unit cells in the donor and acceptor NCs. The (Rn+m+1 )2 distance
dependence for the energy transfer rate, where n and m are the orders of the inter-
acting poles, suggests that higher multipoles are important, even for the shortest
∼7 Å inter-particle separation (R3P/R3PO; R = butyl caps) discussed so far, for
only a small number of unit cells near neighboring NC surfaces. The contributions
to energy transfer from higher multipole interactions are expected to be further
reduced for NCs because the strength of these interactions is weighted by the spa-
tial overlap of electron and hole wavefunctions, which is at a maximum in the NC
center.

Charge Transport Close-packed NC solids are also exciting new materials for
their potentially novel electronic properties. The discrete nature of the NC’s density
of states has generated interest in resonant tunneling of single carriers and in
Coulomb blockade effects in junctions containing a single metal or semiconductor
NC. Building arrays of NCs that demonstrate these novel conduction properties
has raised interest in understanding carrier transport through coupled NCs and in
their potential applications in computational and memory devices (165–167).
At kBT < EC, as required to see Coulomb blockade in single NCs, nonlinear
I-V characteristics have recently been theoretically predicted and experimentally
realized in one- and two-dimensional arrays of metal NCs (168–171) (see further
discussion in Exchange Interactions). As the temperature of the sample is in-
creased, the current threshold disappears and the I-V characteristics become linear
as kBT > EC. These NC solids can be represented as a system of capacitively
597 MURRAY ¥ KAGAN MONODISPERSE
¥ BAWENDI NANOCRYSTAL ASSEMBLIES 597

coupled conductors between which charge carriers tunnel under an applied bias.
The I-V curves show a conduction gap below which little current flows, followed by
onset of carrier tunneling between neighboring NCs. The threshold to conduction
is proportional to the size of the NC array, as the applied voltage must overcome
disorder in the NC energy levels. Above this threshold, the I-V characteristics are
shown to follow the scaling law (169)
µ ¶ζ
V
I∝ −1 15.
VT
Annu. Rev. Mater. Sci. 2000.30:545-610. Downloaded from www.annualreviews.org
Access provided by University of Pennsylvania on 04/13/16. For personal use only.

where VT is the current threshold voltage. In one- and two-dimensional arrays


of Al NCs, consisting of 440 and 38 × 38 NCs, respectively, the exponent ζ is
found to be 1.36 ± 0.01 in one dimension and 1.80 ± 0.16 in two dimensions
(168). Theoretical predictions for arrays of finite size set ζ at 1 and 2, for one- and
two-dimensional arrays, respectively (169). For infinite arrays, calculated values
of ζ are 1 and 5/3 for one- and two-dimensional cases (169).
Kubiak and co-workers have modified the electronic properties of metal NC
arrays by engineering the organic tunnel barriers between neighboring NCs (170).
Exchanging dodecanethiol-capped Au NC arrays with conjugated aryl dithiol and
aryl di-isonitrile molecules serves both to covalently link neighboring NCs and to
provide organic interconnects to control electronic coupling between NCs. Intro-
duction of conjugated linkers may decrease the height of the tunnel barriers, but
in this example it also increases the inter-particle separation, from 13 Å for dode-
canthiol to 17 Å for aryl dithiol. RT conductance measurements show the increase
in NC separation dominates any decrease in barrier height as the conductance
decreases from 133 nS to 78 nS upon exchange.

Photoconductivity The unique, tunable optical properties of semiconductor NCs


present opportunities for use of NC solids as media in photoconductive devices
(172–174) and light-emitting diodes (175–178). Photoconductivity is a well-
known phenomenon in semiconductors and organic solids that has long been used
to probe carrier transport between weakly interacting molecules in otherwise insu-
lating organic solids. The most widely studied semiconductor NC samples to date
are large bandgap materials (Eg > 2 eV). In the dark they have little thermal pop-
ulation of electronic carriers. Semiconductor NCs, like molecules, are also bound
by weak van der Waals interactions in the solid state and, for most organic capping
groups (maintaining inter-particle separations >5 Å), are electrically insulating.
As shown in Figure 26, for CdSe NCs capped with ligands that maintain ∼11 Å
inter-particle separation, the properties of the individual NCs are preserved in the
NC solids. I-V measurements in the dark on these NC solids are not measurable
(on the same instrumentation used for photoconductivity), thus placing a lower
bound on their resistivity of 500 G Ä-cm. Photoconductivity provides a route to
study charge separation and transport in semiconductor NC solids (11, 179). Un-
derstanding charge separation and transport mechanisms and, the inverse, charge
598 MURRAY ¥ KAGAN MONODISPERSE
¥ BAWENDI NANOCRYSTAL ASSEMBLIES 598
Annu. Rev. Mater. Sci. 2000.30:545-610. Downloaded from www.annualreviews.org
Access provided by University of Pennsylvania on 04/13/16. For personal use only.

Figure 30 Model for charge carrier generation in NCs where photoexcitations rapidly
thermalize to their lowest excited state before carrier separation. The |g> represents the
ground state and the |e>is are the excited states of a NC. G is the generation rate of
photoexcitations, kr and knr are the radiative and nonradiative decay rates, respectively. Ä
is the probability of generating free carriers, kF is the field-dependent charge separation,
and R = 1 = Äis the probability of carrier recombination due to their Coulomb attraction.

transport and capture processes, between proximal NCs in the solids is important
for applications of these materials in photoconductive and light-emitting devices.
Photoexcitation of a NC generates an electron-hole pair that is confined to and
delocalized over the volume of the NC. An externally applied electric field is
required to dissociate the exciton, so free electrons and holes can be transported
through the NC solid. Geminate recombination of electron-hole pairs prior to
carrier separation governs the efficiency of photocarrier generation in NC solids,
as is characteristic of other low-mobility solids such as photoconductive polymers
(180), molecular solids (181), intrinsic solids of C60/C70 (182), and some inorganic
solids (183).
Onsager theory for geminate recombination of ions (184), although strictly
solved for classical particles, has successfully described field-assisted charge sep-
aration in many photoconductive solids (185). It models photoconductivity in
solids such as amorphous selenium, where electron-hole pairs are separated, by
the applied electric field, from the excited state to which the photoexcitation was
generated. Onsager theory does not apply to systems such as organic or NC solids,
where internal conversion from higher excited states to the lowest excited state is
very rapid (in ∼ 10−12 to 10−13 s). Figure 30 represents the rapid relaxation pro-
cess and possible fates of photoexcitations in organic and NC solids. The lowest
excited state of most molecules and of NCs is relatively longer lived, ≥1 ns. Car-
rier separation from the lowest excited state of molecules or NCs is an alternate
de-excitation pathway in addition to radiative and nonradiative decay. The charge
generation efficiency (186)
kF (E)
η(E) = 16.
kF (E) + knr + kr
599 MURRAY ¥ KAGAN MONODISPERSE
¥ BAWENDI NANOCRYSTAL ASSEMBLIES 599

where kF (E), the average, field-dependent charge generation rate, competes with
kr and knr, the radiative and nonradiative decay rates (assumed to be weakly field
dependent). The charge separation pathway decreases the PL QY of molecular and
NC solids and can be used as a measure of charge generation efficiency. Onsager’s
model has been modified to represent photoconductivity in organic solids (186).
Although these modified theories conceptually reproduce the dynamics of charge
separation in NCs, they do not account for the appropriate energetics in NCs, as
the exciton binding energy is large compared with kBT.
The applied electric field must overcome the NC size-dependent energy costs to
Annu. Rev. Mater. Sci. 2000.30:545-610. Downloaded from www.annualreviews.org
Access provided by University of Pennsylvania on 04/13/16. For personal use only.

separate the photogenerated charge neutral, quantum confined excitations (11, 179).
The largest energy cost is the binding energy of the exciton, which is inversely
dependent on the square of the NC size. As NCs increase in size, the spatial
overlap of electron and hole wavefunctions decreases, facilitating exciton dis-
sociation. To separate the charge neutral photoexcitation on a single NC and
generate two charged NCs, one with an electron and the other hole, the charging
energy of the NCs must be supplied by the electric field. As described above
for metal NCs, the charging energy of the NC is also inversely dependent on
size. Finally, for two neighboring, oppositely charged NCs, a Coulomb attrac-
tion still must be overcome. The Coulomb energy is inversely dependent on the
distance between charges, which for NC solids depends on NC size as it defines
the electron-to-hole site-to-site distance. The efficiency of charge separation also
depends on the surface passivation for the NCs. Decreasing the distance between
NCs, for example, by shortening the length of the alkyl chain on R3P/R3PO caps
from octyl to butyl, increases the probability of carrier tunneling. The degree to
which the organic ligands electronically passivate the NC surfaces may also affect
charge separation because surface traps act as stepping-stones to facilitate charge
separation and transport. Recently a resonant tunneling model was developed to
qualitatively represent size- and surface-dependent photoconductivity in NC solids
(179).
Photocarrier generation from NC excitations can be seen in the spectral response
for a representative size series of CdSe NC solids (Figure 31) (11, 179). The
spectral response of the photocurrent maps the discrete, size-dependent electronic
transitions of the NCs in the solids as measured in absorption. Correlation between
the spectral response and absorption has also been observed in InP NC solids (159).
The carrier generation efficiency for the NC solids is independent of the excitation
energy. An increase in the magnitude of the photocurrent as the excitation energy
is tuned to the blue would be expected if hot carriers were separated before being
thermalized. The rapid relaxation of excitations (i.e. no phonon bottleneck) can
similarly be seen in PL measurements for NCs, as only emission from the band-
edge is observed. The relative decrease in the magnitude of the photocurrent for
bluer relative to redder excitation energies is consistent with a measured decrease
in the PL QY for dispersed NC samples.
The dynamics represented in Figure 30 are also consistent with temperature-
dependent photoconductivity measurements. The photocurrent strongly decreases
600 MURRAY ¥ KAGAN MONODISPERSE
¥ BAWENDI NANOCRYSTAL ASSEMBLIES 600
Annu. Rev. Mater. Sci. 2000.30:545-610. Downloaded from www.annualreviews.org
Access provided by University of Pennsylvania on 04/13/16. For personal use only.

Figure 31 Photocurrent spectral responses (circles) map the absorption spectra (solid
lines) of the NCs in the solids for a size series of solids prepared from CdSe NC samples
(A) 35, (B) 41.5, (C ) 49.5, and (D) 60 Å in diameter.

with increasing temperature. Equation 16 shows that the efficiency of charge gen-
eration is inversely proportional to the radiative and nonradiative decay rates. Cal-
culations of the temperature-dependent PL QY, from absorption and PL measure-
ments of the NC solids, show an exponential decrease with increasing temperature
as the nonradiative rate of decay increases.
The measured photocurrent is linear with increasing excitation intensity. Photon
photons
fluxes of 1016 cm2 · sec impinge ∼10
12 photons/s on the active area of the NC solids

[L = 20 µm, W = 800 µm] and generate photocurrents of ∼30 pA at applied


voltages of ± 500 V. The thickness of the NC solids is on the order of the pen-
etration distance of the light. Therefore photoexcitations are created through the
601 MURRAY ¥ KAGAN MONODISPERSE
¥ BAWENDI NANOCRYSTAL ASSEMBLIES 601

thickness of the NC solids. For an optical density of 0.1, the external carrier
generation efficiency for the NC solids is ∼5 × 10−4.
Photoconductivity has also been studied in NC-doped conjugated polymers
(187, 188). While NC doped polymers are not the focus of this review, at high
NC concentrations, TEM images show some regions of nearly close-packed NCs.
There is also some commonality between the physics of charge separation in close-
packed NC solids and in NC-doped conjugated polymers. Onsager’s model for
geminate recombination has also been applied to describe photoconductive dis-
charge measurements in CdS-doped poly(vinyl carbazole) (PVK) (187). Because
Annu. Rev. Mater. Sci. 2000.30:545-610. Downloaded from www.annualreviews.org
Access provided by University of Pennsylvania on 04/13/16. For personal use only.

many of the photoconductive and light-emitting devices have been prepared as


close-packed NC layers sandwiched between polymeric layers or as NC-polymer
mixtures, understanding the physics of charge transport at the NC-polymer inter-
face is also important.

Strong Coupling/Exchange Interactions


Semiconductor NC Solids As the distance between atoms, molecules, or NCs
is reduced to separations ≤5 Å, exchange interactions become significant. In
semiconductor NC samples, the onset of this stronger interaction has been observed
in pyridine-capped CdSe NCs. For larger NC sizes, for example CdSe NCs ≥ 25 Å
in diameter, pyridine-capped NCs (shown in Figure 11) still maintain an ∼7 Å
inter-particle separation, and the optical absorption spectra appear unchanged.
Because pyridine is a volatile solvent, applying vacuum and gently heating the NC
solid removes the pyridine cap, collapsing the inter-particle separation, and upon
further heating sinters the NCs in the solid (10). As the distance between NCs is
reduced, from 7 to 0 Å, the optical absorption spectrum red shifts, approaching
the absorption spectrum for bulk CdSe (10, 11). The strength of the interaction is
such that, as in bulk, inorganic, or molecular solids, the time constant for energy
transfer is comparable to that for the absorption process. Electronic excitations are
delocalized over many NCs or, for sintered solids, over many unit cells as in bulk
CdSe. Electronic states, and their associated wavefunctions, are now described by
linear combinations of multiple NCs in the solids or for sintered solids by Bloch
functions.
Woggon and co-workers recently showed that the optical spectra of close-
packed NC solids, prepared from the smallest, ∼16 Å diameter CdSe NC samples
with pyridine caps, look like that for bulk CdSe (189). As the size of the NCs
decreases, the electronic wavefunctions from the NC states spill further outside the
volume of the NCs. In the smallest species, the extension of the electronic wave-
functions toward neighboring NCs may be large enough for exchange interactions
to become significant and electronic excitations to become delocalized.

Metal NC Solids Close-packing metal NCs into solids also gives rise to unique,
tunable optical and electronic properties in the solid state. As in the semiconductor
systems, the organic ligands coordinating the metal NCs in the solids mediate
602 MURRAY ¥ KAGAN MONODISPERSE
¥ BAWENDI NANOCRYSTAL ASSEMBLIES 602

electronic coupling between NCs and therefore the properties of the solids. For
example, the energy of the surface plasmon resonance of individual metal NCs
is affected by the dielectric constant of the matrix. As metal NCs are placed in
proximity in the solids, the surface plasmon resonance shifts to lower energy as
the average dielectric constant of the surrounding medium increases (90, 190).
The surface plasmon resonance, particularly for the noble metal NCs that are a
focus of this review, are strong, originating from high oscillator strength transition
dipole moments. Coupling between transition dipoles or higher multipoles, as for
semiconductor NC solids described above, enhances the nonlinear optical response
Annu. Rev. Mater. Sci. 2000.30:545-610. Downloaded from www.annualreviews.org
Access provided by University of Pennsylvania on 04/13/16. For personal use only.

of metal NC solids. At inter-particle distances >5Å , maintained by most organic


capping groups, electronic excitations and carriers are localized in individual NCs,
and the solids are electrically insulating (see Charge Transport) with a Coulomb
bandgap.
At inter-particle distances ≤5 Å , exchange interactions and electronic delocal-
ization become significant. Heath and co-workers have shown an insulator-to-
metal transition as the distance between metal NCs decreases by preparing and
compressing monolayers of alkanethiol-capped Ag NCs on a Langmuir trough
(190–195). The strength of electronic coupling between NCs and its influence on
the properties of metal NC solids is described by the cartoon in Figure 32 (top).
As the trough barriers are closed, the inter-particle separation decreases, such that
the distance between NC centers (D) becomes comparable to twice the NC radii
(2R). The coupling between NCs increases and the energy bands of the NC solids
increase in width, decreasing the Coulomb bandgap, with decreasing D/2R. When
D/2R ∼ 1.2, exchange interactions give rise to the insulator-to-metal transition
and the Coulomb bandgap disappears. The onset of electronic delocalization is
observed by dramatic changes in the optical and electronic properties of the metal
NC monolayers.
The insulator-to-metal transition was first observed in the linear and nonlinear
optical absorption and reflection of these monolayers (191). As the distance be-
tween Ag NCs on the liquid subphase decreases to form a continuous monolayer,
the surface plasmon resonance continually decreases in energy and the absolute
reflectivity increases, as described by weak coupling of electromagnetic fields.
Just before the monolayer collapses to a bilayer structure, the distance between
surfaces for NCs capped with sufficiently short chain length ligands reaches ∼5 Å,
the typical distance for the onset of exchange interactions. At these short inter-
particle separations, electronic wavefunctions are delocalized over many NCs in
the monolayers and the monolayer behaves as a metal. The reflectivity of the
monolayer drops, and the absorption increases to levels characteristic of thin, con-
tinuous metal films. The linewidth of the measured surface plasmon resonance
decreases as delocalization of electronic wavefunctions increases the scattering
length and time of electronic carriers. These changes in the linear optical proper-
ties are accompanied by an exponential increase and then a sharp discontinuity in
χ (2) measured from the second harmonic generation signal. The observed changes
in the optical spectra from the insulator-to-metal transition are reversible. Opening
Annu. Rev. Mater. Sci. 2000.30:545-610. Downloaded from www.annualreviews.org
Access provided by University of Pennsylvania on 04/13/16. For personal use only.

Figure 32 (Top) Cartoon of inter-particle interactions. At large inter-particle distance (D), NCs
are electrically isolated and the superlattice is an insulator with a Coulomb bandgap. As the inter-
particle distance decreases, exchange interactions become significant and the localized electronic
wavefunctions of the individual NCs spread out over multiple NCs in the superlattice. In metal NC
superlattices exchange interactions lead to an insulator-to-metal transition. (Bottom) (a) dl/dV
or normalized density of states (NDOS) versus applied V for (a) decanethiol-capped Ag NCs
maintaining D ∼ 1.2 nm and (b) hexanethiol-capped Ag NCs providing D ∼ 0.5 nm (195).
603 MURRAY ¥ KAGAN MONODISPERSE
¥ BAWENDI NANOCRYSTAL ASSEMBLIES 603

the barriers used to compress the monolayer of Ag NCs increases the inter-particle
separation, and the optical properties of the monolayer is again characteristic of
an insulator. The second harmonic generation signal and the surface plasmon
linewidth are modeled by accounting for quantum mechanical coupling of elec-
tronic wavefunctions between neighboring NCs (192, 193).
These changes in optical properties of metal NC monolayers are consistent
with more recent electrical measurements by impedance spectroscopy (194) and
scanning tunneling microscopy (STM) (195). STM measurements of 26 Å Ag
NC monolayers are shown in Figure 32 for decanethiol caps, which maintain
Annu. Rev. Mater. Sci. 2000.30:545-610. Downloaded from www.annualreviews.org
Access provided by University of Pennsylvania on 04/13/16. For personal use only.

12 Å inter-particle separations, and for hexanethiol caps, which maintain 5 Å

separations. The finite size of the 26 Å Ag NCs gives rise to a 0.34 eV charging
energy, EC. At the larger inter-particle separations provided by the decanethiol
caps, the Ag NC monolayers are electrically insulating and have a classic Coulomb
bandgap and staircase that open up as the temperature of the sample is reduced from
90 to 20 K. The normalized conductances are related to the normalized density of
states (NDOS), for data collected at sufficiently low voltage biases. The insulating
character of the metal NC solids with decanethiol caps is seen in Figure 32a, by the
vanishing density of states at zero bias. As the temperature decreases, the bandgap
opens further as thermally activated phonons become unavailable to assist carrier
hopping between the electrically isolated NCs.
As the distance between NCs decreases, as shown for hexanethiol caps, the
Ag NC monolayers are metallic. The non-zero and temperature-independent nor-
malized density of states, shown at RT and 20 K in Figure 32b, is characteristic
of a metal and is comparable to evaporated Ag thin films. The characteristic fea-
tures arising from the charging energy of the finite size, individual Ag NC building
blocks of these monolayers have disappeared. The decrease in kinetic energy as the
electronic wavefunctions become delocalized over many NCs in the monolayers
overcomes the energy cost to charge the Ag NCs.
Recently Pb NCs and their solids have been prepared to study Josephson
exchange coupling, which would lead to delocalization of the superconducting
Cooper pairs over multiple NCs in the solids (196). If a material is a type I su-
perconductor (e.g. Pb), below Tc it will generate a supercurrent of Cooper pairs
to exclude an applied magnetic field, a phenomenon known as the Meisner effect.
SQUID magnetometry is used to probe the superconducting state; the materials
response to oppose an applied magnetic field, known as diamagnetism, results in
a measured negative magnetic susceptibility. The applied field physically pene-
trates a superconductor to a characteristic skin depth at which the opposing field
is equal in magnitude. If the size of the superconducting material is smaller than
the skin depth, the applied magnetic field dominates, which returns the material to
its normal conducting state. As shown by Heath and co-workers, by preparing Pb
NCs ranging in size from 150 to 700 Å in diameter, Pb NCs smaller than 450 Å su-
perconduct but are too small to show a Meisner effect, and those larger than 450 Å
superconduct and behave like bulk Pb. Josephson exchange coupling is studied
604 MURRAY ¥ KAGAN MONODISPERSE
¥ BAWENDI NANOCRYSTAL ASSEMBLIES 604
Annu. Rev. Mater. Sci. 2000.30:545-610. Downloaded from www.annualreviews.org
Access provided by University of Pennsylvania on 04/13/16. For personal use only.

Figure 33 Magnetic susceptibility versus temperature at 400 Gauss for Pb QD assemblies


separated by dodecane- (C12), octane- (C8), and hexane- (C6)carboxylates (196).

by choosing small 200 Å Pb NCs. As shown in Figure 33, for Pb NCs capped

with long organic capping groups such as dodecanecarboxylate, the Pb NCs are
electronically isolated and no diamagnetic response is observed as the material is
in its normal conducting state. As the length of the organic capping group is de-
creased to octanecarboxylates and hexanecarboxylates, the diamagnetic response
of the Pb NC solids increases. This suggests that the superconducting length scale
is increased beyond the skin depth, i.e. Josephson exchange coupling results in
delocalization of Cooper pairs in the Pb NC solids.
605 MURRAY ¥ KAGAN MONODISPERSE
¥ BAWENDI NANOCRYSTAL ASSEMBLIES 605

CONCLUSION

Methods have been established for the preparation of monodisperse NC sam-


ples for a variety of semiconducting and metallic materials. These samples are
monodisperse in size, to within the limits of atomic roughness, and shape, and have
well-formed crystalline cores and controlled surface chemistry. Increasingly, a di-
verse array of chemical and structural probes is being employed to characterize
NC samples. When the results of these independent techniques are united through
simple computer simulations, a self-consistent model of NC structure emerges.
Annu. Rev. Mater. Sci. 2000.30:545-610. Downloaded from www.annualreviews.org
Access provided by University of Pennsylvania on 04/13/16. For personal use only.

Preparation and structural characterization of a homologous size series of such


NC samples enables the systematic study of the electronic and optical proper-
ties inherent to nanometer size materials. In nanometer size semiconductors and
metals, three-dimensional quantum confinement effects collapse the continuous
density of states of the bulk solid into the discrete electronic states of the NC. These
effects lead to unique electronic and optical properties that may be exploited in
future single NC devices.
The rational assembly of solids from nanometer scale building blocks presents
opportunities to tailor the properties of solids by choosing the NCs, by size or com-
position, or the interactions between NCs, by the length and chemical functionality
of the organic inter-layer. By tailoring the composition of solvents, glassy and or-
dered three-dimensional solids may be deposited. Development of a well-defined
structural picture, using a variety of structural probes, is required to understand
collective interactions in the solid state. Interactions between neighboring NCs
can be weak, such as dipolar coupling giving rise to energy transfer or charge
tunneling leading to dark and photoconductivity. As the strength of inter-particle
interactions increases, exchange interactions may lead to semiconducting, metal-
lic, or superconducting NC solids.

ACKNOWLEDGMENTS
The authors thank J R Heath, D J Norris, F Mikulec, B A Korgel, D Fitzmaurice,
AV Kadavanich, OI Micic, and AJ Nozik for contributing figures on their work to
make this review more complete.

Visit the Annual Reviews home page at www.AnnualReviews.org

LITERATURE CITED
1. Kubo R, Kawabata A, Kobayashi S. 1984. 5. Gaponenko SV. 1998. Optical Properties of
Annu. Rev. Mater. Sci. 14:49–66 Semiconductor Nanocrystals. Cambridge,
2. Hayashi C, Uyeda R, Tasaki A, eds. 1997. UK: Cambridge Univ. Press
Ultra-Fine Particles. Westwood, NJ: Noyes 6. Woggon U. 1997. Optical Properties
3. Perenboom JAAJ, Wyder P, Meier F. 1981. of Semiconductor Quantum Dots. Berlin:
Phys. Rep. 78:173–292 Springer-Verlag
4. Nagaev EL. 1992. Phys. Rep. 222:199–307 7. Kastner MA. 1993. Phys. Today 46:24–31
606 MURRAY ¥ KAGAN MONODISPERSE
¥ BAWENDI NANOCRYSTAL ASSEMBLIES 606

8. Alivisatos AP. 1996. Science 271:933–37 26. Brust M, Walker M, Bethell D, Schiffrin
9. Brus L. 1998. J. Phys. Chem. Solids DJ, Whyman R. 1994. J. Chem. Soc. Chem.
59:459–65 Commun. 801–2
10. Murray CB. 1995. Synthesis and charac- 27. Leff DV, Brandt L, Heath JR. 1996. Lang-
terization of II-VI quantum dots and their muir 12:4723–30
assembly into 3D quantum dot superlat- 28. Murthy S, Bigioni TP, Wang ZL, Khoury
tices. PhD thesis. Massachusetts Institute JT, Whetten RL. 1997. Mater. Lett. 30:321–
of Technology, Cambridge, MA. 164 pp. 25
11. Kagan CR. 1996. The electronic and op- 29. Taleb A, Petit C, Pileni MP. 1997. Chem.
tical properties of close packed cadmium Mater. 9:950–59
selenide quantum dot solids. PhD the- 30. Motte L, Billoudet F, Lacaze E, Pileni MP.
Annu. Rev. Mater. Sci. 2000.30:545-610. Downloaded from www.annualreviews.org
Access provided by University of Pennsylvania on 04/13/16. For personal use only.

sis. Massachusetts Institute of Technology, 1996. Adv. Mater. 8:1018–20


Cambridge, MA. 168 pp. 31. Petit C, Taleb A, Pileni MP. 1999. J. Phys.
12. Overbeek JThG. 1982. Adv. Coll. Int. Sci. Chem. B 103:1805–10
15:251–77 32. Sun S, Murray CB. 1999. J. Appl. Phys
13. La Mer VK, Dinegar RH. 1950. J. Am. 85:4325–30
Chem. Soc. 72:4847–54 33. Bradley JS, Hill EW, Klein C, Chaudret
14. Reiss H. 1951. J. Chem. Phys. 19:482 B, Duteil A. 1993. Chem. Mater. 5:254–
15. De Smet Y, Deriemaeker L, Finsy R. 1997. 56
Langmuir 13:6884–88 34. Kortan AR, Hull R, Opila RL, Bawendi
16. Grätz H. 1997. Scr. Mater. 37:9–16 MG. Steigerwald ML, et al. 1990. J. Am.
17. Murray CB, Norris DJ, Bawendi MG. Chem. Soc. 112:1327–32
1993. J. Am. Chem. Soc. 115:8706–15 35. Hines MA, Guyot-Sionnest P. 1996. J.
18. Mikulec F. 1999. Semiconductor nano- Phys. Chem. 100:468–71
crystal colloids: manganese doped cad- 36. Dabbousi BO, Rodriguez-Viejo J, Mikulec
mium selenide, (core) shell composites for FV, Heine JR, Mattoussi H, et al. 1997. J.
biological labeling, and highly fluores- Phys. Chem. B 101:9463–75
cent cadmium telluride. PhD. thesis. Mas- 37. Danek M, Jensen KF, Murray KF, Mur-
sachusetts Institute of Technology, Cam- ray CB, Bawendi MG. 1996. Chem. Mater.
bridge, MA. 8:173–80
19. Hines MA, Guyot-Sionnest P. 1998. J. 38. Peng X, Schlamp MC, Kadavanich AV,
Phys. Chem. B 102:3655–57 Alivisatos AP. 1997. J. Am. Chem. Soc.
20. Micic OI, Sprague JR, Curtis CJ, Jones 119:7019–79
KM, Nozik AJ. 1994. J. Phys. Chem. 98: 39. Mews A, Eychmüller A, Giersig M,
4966–69 Schooss D, Weller H. 1994. J. Phys. Chem.
21. Guzelian AA, Bowen Katari JE, Kada- 98:934–41
vanich AV, Banin U, Hamad K, et al. 1996. 40. Kershaw SV, Burt M, Harrison M, Rogach
J. Phys. Chem. 100:7212–19 A, Weller H, Eychmüller A. 1999. Appl.
22. Guzelian AA, Banin U, Kadavanich AV, Phys. Lett. 75:2694–96
Peng X, Alivisatos AP. 1996. Appl. Phys. 41. Mews A, Kadavanich AV, Banin U,
Lett. 69:1432–34 Alivisatos AP. 1996. Phys. Rev. B
23. Turkevich J, Stevenson PC, Hillier J. 1951. 53:R13242–45
Disc. Faraday Soc. 11:55 42. Peng X, Wickham J, Alivisatos AP. 1998.
24. Schmid G. 1992. Chem. Rev. 92:1709–27 J. Am. Chem. Soc. 120:5343–44
25. Schmid G. 1994. Clusters and Col- 43. Steigerwald ML, Alivisatos AP, Gibson
loids: From Theory to Applications. New JM, Harris TD, Kortan AR, et al. 1988. J.
York:Wiley & Sons Am. Chem. Soc. 110:3046–51
607 MURRAY ¥ KAGAN MONODISPERSE
¥ BAWENDI NANOCRYSTAL ASSEMBLIES 607

44. Kuno M, Lee JK, Dabbousi BO, Mikulec 62. Shiang JJ, Kadavanich AV, Grubbs RK,
FV, Bawendi MG. 1997. J. Chem. Phys. Alivisatos AP. 1995. J. Phys. Chem.
106:9869–82 99:17417–22
45. Murray CB, Kagan CR, Bawendi MG. 63. Hÿtch MJ, Gandais M. 1995. Philos. Mag.
1995. Science 270:1335–38 A 72:619–34
46. Evans RML, Fairhurst DJ, Poon WCK. 64. Malm JO, O’Keefe MA. 1997. Ultrami-
1998. Phys. Rev. Lett. 81:1326–29 croscopy 68:13–23
47. Chemseddine A, Weller H. 1993. Ber. 65. Dinega DP, Bawendi MG. 1999. Angew.
Bunsen-Ges, Phys. Chem. 97:636–37 Chem. 38:1788
Annu. Rev. Mater. Sci. 2000.30:545-610. Downloaded from www.annualreviews.org

48. Yamaki M, Higo J, Nagayama K. 1995. 66. Kagan CR, Murray CB, Bawendi MG.
Access provided by University of Pennsylvania on 04/13/16. For personal use only.

Langmuir 11:2975–78 1995. MRS Symp. Proc. 358:219–24


49. Ohara PC, Leff DV, Heath JR, Gelbart 67. Kagan CR, Murray CB, Bawendi MG.
WM. 1995. Phys. Rev. Lett. 75:3466– 1996. Phys. Rev. B 54:8633–43
69 68. Korgel BA, Fitzmaurice D. 1999. Phys.
50. Doome RJ, Fonseca A, Richter H, Nagy Rev. B 59:14191–20
JB, Thiry PA, Lucas AA. 1997. J. Phys. 69. Korgel BA, Fitzmaurice D. 1998. Phys.
Chem. Sol. 58:1839–43 Rev. Lett. 80:3531–34
51. Fischer Ch-H, Weller H, Fojtik A, Lume- 70. Glatter O, Kratky O. 1982. Small Angle X-
Pereira C, Janata E, Henglein A. 1986. Ber. ray Scattering. New York: Academic
Bunsen-Ges. Phys. Chem. 90:46–49 71. Guinier A, Fournet G. 1955. Small Angle
52. Wilcoxon JP, Martin JE, Parsapour F, Scattering of X-Rays. New York: Wiley
Wiedenman B, Kelley DB. 1998. J. Chem. &Sons
Phys. 108:9137–43 72. Caponetti E, Floriano MA, Di Dio E, Triolo
53. Alvarez MM, Kjoury JT, Schaaff TG, R. 1993. J. Appl. Cryst. 26:612–25
Shafigullin MN, Vezmar I, Whetten RL. 73. Allen AJ, Berk NF. 1994. J. Appl. Cryst.
1997. J. Phys. Chem. B 101:3706–12 27:878–91
54. Sachleben JR, Colvin V, Emsley L, Wooten 74. Bawendi MG, Kortan AR, Steigerwald
EW, Alivisatos AP. 1998. J. Phys. Chem. B ML, Brus LE. 1989. J. Chem. Phys.
102:10117–28 91:7282–90
55. Majetich SA, Carter AC, Belot J, McCul- 75. Hall BD, Monot R. 1991. Comput. Phys.
lough RD. 1994. J. Phys. Chem. 98:13705– 5:414–17
10 76. Boon KT, 1986. EXAFS: Basic Principles
56. Bowen Katari JE, Colvin VL, Alivisatos and Data Analysis. Chpt. 1–3. New York:
AP. 1994. J. Phys. Chem. 98:4109–17 Springer-Verlag
57. Wang ZL, Harfenist SA, Whetten RL, 77. Marcus MA, Brus LE, Murray CB,
Bentley J, Evands ND. 1998. J. Phys. Bawendi MG, Prasad A, Alivisatos AP.
Chem. B 102:3068–72 1992. Nanostruct. Mater. 323:1
58. Becerra LR, Murray CB, Griffin RG, 78. Bradley JS, Via GH, Bonneviot L, Hill EW.
Bawendi MG. 1994. J. Chem. Phys. 1996. Chem. Mater. 8:1895–903
100:3297–300 79. Herron N, Calabrese JC, Farneth WE,
59. Tomaselli M, Yarger JL, Bruchez M Jr, Wang Y. 1993. Science 259:1426–28
Havlin RH, deGraw D, et al. 1999. J. Chem. 80. Vossemeyer T, Reck G, Katsikas L, Haupt
Phys. 110:8861–64 ETK, Schulz B, Weller H. 1995. Science
60. Doraiswamy N, Marks LD. 1995. Philos. 267:1476
Mag. B 71:291–310 81. Rockenberger J, Tröger L, Kornowski A,
61. Marks LD. 1994. Rep. Prog. Phys. 57:603– Vossmeyer T, Eychmüller A, et al. 1997. J.
49 Phys. Chem. B 101:2691–701
608 MURRAY ¥ KAGAN MONODISPERSE
¥ BAWENDI NANOCRYSTAL ASSEMBLIES 608

82. Efros AlL, Efros AL. 1982. Sov. Phys. 103. Bentzon MD, van Wonterghem J, Morup
Semicond. 16:772 S, Thölén A. 1989. Philos. Mag. B
83. Brus LE. 1984. J. Chem. Phys. 80:4402 60:169–78
84. Norris DJ, Efros Al-L, Rosen M, Bawendi 104. Bentzon MD, Thölén AR. 1991. Ultrami-
MG. 1996. Phys. Rev. B 53:16347–54 croscopy 38:105–15
85. Norris DJ, Bawendi MG. 1996. Phys. Rev. 105. Vossmeyer T, Katsikas L, Gersig M,
B 53:16338–46 Popovic IG, Diesner K, et al. 1994. J.
86. Banin U, Lee CJ, Guzelian AA, Kada- Phys. Chem. 98:7665–73
vanich AV, Alivisatos AP, et al. 1998. J. 106. Ptatschek V, Schreder B, Herz K, Hilbert
Chem. Phys. 109:2306–9 U, Ossau W, et al. 1997. J. Phys. Chem.
87. Banin U, Cerullo G, Guzelian AA, B 101:8898–906
Annu. Rev. Mater. Sci. 2000.30:545-610. Downloaded from www.annualreviews.org
Access provided by University of Pennsylvania on 04/13/16. For personal use only.

Bardeen CJ, Alivisatos, Shank CV. 1997. 107. Lowen H. 1994. Melting, Freezing, and
Phys. Rev. B 55:7059–67 Colloidal Suspensions, pp. 249–324. New
88. Empedocles SA, Bawendi MG. 1999. York, NY: Elsevier
Acc. Chem. Res. 32:389–96 108. Gacoin T, Malier L, Boilot JP. 1997. J.
89. Ashoori RC. 1996. Nature 379:413–19 Mater. Chem. 7:859–60
90. Kreibeg U, Vollmer M. 1995. Optical 109. Pusey PN, van Megen W. 1994. J. Phys.
Properties of Metal Clusters, Vol. 25. Condens. Matter 6:A29–36
New York: Springer Verlag. 532 pp. 110. North AN. 1991. In Fine Ceramic Pow-
91. Grabert H. 1992. In Single Charge Tun- ders, ed. R Freer, JL Woodhead, pp. 1–
neling, ed. MH Devoret, H Grabert, pp. 151. Shelton, UK: Institute of Ceramics
1–335. New York: Plenum 111. Ottewill RH. 1991. J. Appl. Cryst.
92. Klein DL, McEuen PL, Katari JEB, Roth 24:436–43
R, Alivisatos AP. 1996. Appl. Phys. Lett. 112. Goodwin JW, Buscall R. 1995. Colloidal
68:2574–76 Polymer Particles. Boston: Academic.
93. Black CT, Ralph DC, Tinkham M. 1996. 352 pp.
Phys. Rev. Lett. 76:688–91 113. Cusack NE. 1987. The Physics of Struc-
94. Ingram RS, Hostetler MJ, Murray RW, turally Disordered Matter. Philadelphia:
Schaaff TG, Khoury JT, et al. 1997. J. Am. Hilger. 402 pp.
Chem. Soc. 119:9279–80 114. Schaaff TG, Shafigullin MN, Khoury JT,
95. Andres RP, Bein T, Dorogi M, Feng Vezmar I, Whetten RL, et al. 1997. J.
S, Henderson JI, et al. 1996. Science Phys. Chem. B 101:7885–91
272:1323–28 115. Kiely CJ, Fink J, Brust M, Bethell D,
96. Black CT. 1996. Tunneling spectroscopy Schiffrin DJ. 1998. Nature 396:444–46
of nanometer-scale metal particles. PhD 116. Gregory J. 1996. Particle aggregation:
thesis. Harvard Univ., Cambridge, MA. modeling and measurement. In Nanopar-
189 pp. ticles in Solids and Solutions, ed. JH
97. Fulton TA, Dolan GJ. 1987. Phys. Rev. Fendler, I Dékány, pp. 203–55. Nether-
Lett. 59:109 lands: Kluwer
98. Sato T, Ahmed H, Brown D, Johnson 117. Zhu PW, White JW. 1996. J. Chem. Phys.
BFG. 1997. J. Appl. Phys. 82:696–701 104:9169–73
99. Sanders JV. 1967. Acta Cryst. A24:427 118. Connolly S, Fullam S, Korgel B, Fitz-
100. Pieranski P. 1983. Cont. Phys. 25:25 maurice D. 1998. J. Am. Chem. Soc.
101. Van Blaarderen A, Ruel R. Wiltzius P. 120:2969–70
1997. Nature 385:321–24 119. Korgel BA, Zaccheroni N, Fitzmaurice D.
102. Murray CA, Grier DG. 1995. Am. Sci. 1999. J. Am. Chem. Soc. 121:3533–34
83:238 120. Deleted in proof
609 MURRAY ¥ KAGAN MONODISPERSE
¥ BAWENDI NANOCRYSTAL ASSEMBLIES 609

121. Burnside SD, Shklover V, Barbé C, Comte 140. Niemeyer CM. 1999. Appl. Phys. A
P, Arendse F, et al. 1998. Chem. Mater. 68:119–24
10:2419–25 141. Mirkin CA, Letsinger RL, Mucic RC,
122. Whetten RL, Shafigullin MN, Khoury JT, Storhoff JJ. 1996. Nature 382:607–9
Schaaff TG, Vezmar I, et al. 1999. Acc. 142. Alivisatos AP, Johnsson KP, Peng X, Wil-
Chem. Res. 32:397–406 son TE, Loweth CJ, et al. 1996. Nature
123. Taleb A, Petit C, Pileni MP. 1998. J. Phys. 609–11
Chem. B 102:2214–20 143. Esaki L, Tsu R. 1970. IBM J. Res. De-
124. Petit C, Taleb A, Pileni MP. 1999. J. Phys. velop. 14:61–65
Chem. B 103:1805–10 144. Heitmann D, Kotthaus JP. 1993. Phys. To-
125. Yin JS, Wang ZL. 1997. Phys. Rev. Lett. day 46:56–63
Annu. Rev. Mater. Sci. 2000.30:545-610. Downloaded from www.annualreviews.org
Access provided by University of Pennsylvania on 04/13/16. For personal use only.

79:2570–73 145. Kempa K, Broido DA, Pakshi P. 1991.


126. Li M, Schnablegger H, Mann S. 1999. Na- Phys. Rev. B 43:9343–45
ture 402:393–95 146. Kotlyar R, Stafford CA, Das Sarma S.
127. Frankel F, Whiteside GM. 1997. On the 1998. Phys. Rev. B 58:3989–4013
Surface of Things. San Francisco: Chron- 147. Govorov AO, Chaplik AV. 1994. J. Phys.
icle Books. p. 80 Condens. Matter 6:6507–14
128. Burton WK, Cabrera N, Frank FC. 1951. 148. Chaplik AV, Ioriatti L. 1992. Surf. Sci.
Philos. Trans R. Soc. London Ser. A 263:354–58
243:299–358 149. Gudmundsson V. 1998. Phys. Rev. B
129. Levi AC, Kotrla M. 1997. J. Phys. Con- 57:3989–93
dens. Matter 9:299–344 150. Lloyd S. 1993. Science 261:1569–71
130. Tersoff J, Teichert C, Lagally MG. 1996. 151. Lent CS, Tougaw PD, Porod W. 1993.
Phys. Rev. Lett. 76:1675–78 Appl. Phys. Lett. 62:714–16
131. Budai JD, White CW, Withrow SP, 152. Kagan CR, Murray CB, Nirmal M,
Chisholm MF, Zhu J, Zuhr RA. 1997. Na- Bawendi MG. 1996. Phys. Rev. Lett.
ture 390:384–86 76:1517–20
132. Blanton SA, Leheny RL, Hines MA, 153. Agranovich VM, Galanin MD. 1982.
Guyot-Sionnest P. 1997. Phys. Rev. Lett. Electronic Excitation Energy Transfer in
79:865–68 Condensed Matter. New York: North-
133. Harfenist SA, Wang ZL, Alvarez MM, Holland
Vezmar I, Whetten RL. 1996. J. Phys. 154. Förster Th. 1960. In Comparative Effects
Chem. 100:13904–10 of Radiation, ed. M Burton, JS Kirby-
134. Safran SA, Clark NA. 1987. Physics Smith, JL Magee, pp. 300–41. New York:
of Complex and Supermolecular Fluids. Wiley
New York: Wiley & Sons 155. Dexter DL. 1953. J. Chem. Phys. 21:836–
135. Eldridge MD, Madden PA, Frenkel D. 50
1993. Nature 365:35–37 156. Windsor MW. 1965. Luminescence and
136. Bartlett P, Ottewill RH, Pusey PN. 1990. energy transfer. In Physics and Chemistry
J. Chem. Phys. 93:1299–312 of the Organic Solid State, ed. D Fox,
137. Vossmeyer T, DeIonno E, Heath JR. 1997. 2:343–31. New York: Interscience
Angew. Chem. Int. Ed. Engl. 36:1080–83 157. Lamola AA. 1969. Electronic energy
138. Vossmeyer T, Jia S, Delonno E, Diehl transfer in solution: theory and ap-
MR, Kim S-H, et al. 1998. J. Appl. Phys. plications. In Energy Transfer and Or-
84:3664–70 ganic Photochemistry, ed. AA Lam-
139. Vlasov YA, Nan Y, Norris DJ. 1999. Adv. ola, NJ Truro, 14:17–132. New York:
Mater. 11:165–69 Interscience
610 MURRAY ¥ KAGAN MONODISPERSE
¥ BAWENDI NANOCRYSTAL ASSEMBLIES 610

158. Liu G, Guillet JE. 1990. Macromolecules 1997. J. Appl. Phys. 82:5837–42
23:1388 179. Leatherdale CA, Kagan CR, Morgan NY,
159. Micic OI, Jones KM, Cahill A, Nozik AJ. Empedocles SA, Kastner MA, Bawendi
1998. J. Phys. Chem. B 102:9791–96 MG, 2000. Phys. Rev. B. Submitted
160. Vossmeyer T, Reck G, Katsikas L, Haupt 180. Gill WD. 1976. In Photoconductivity and
ETK, Schulz B, Weller H. 1995. Science Related Phenomena. ed. J Mort, DM Pai,
267:1476–79 p. 303. New York: Elsevier
161. Pullerits T, Freiberg A. 1991. Chem. Phys. 181. Inokuchi H, Maruyama K. 1976. In Pho-
149:409 toconductivity and Related Phenomena.
162. Somsen OJG, van Mourik F, van Gron- ed. J Mort, DM Pai, p. 155. New York:
delle R, Valkunas L. 1994. Biophys. J. Elsevier
Annu. Rev. Mater. Sci. 2000.30:545-610. Downloaded from www.annualreviews.org
Access provided by University of Pennsylvania on 04/13/16. For personal use only.

66:1580 182. Mort J, Machonkin M, Ziolo R, Chen I.


163. Takagahara T. 1993. Optoelect. Dev. Tech. 1992. Appl. Phys. Lett. 61:1829
8:545–55 183. Pai DM, Enck RC. 1975. Phys. Rev. B
164. Kayanuma Y. 1993. J. Phys. Soc. Jpn. 11:5163
62:346–56 184. Onsager L. 1938. Phys. Rev. 54:554–57
165. Heath JR, Kuekes PJ, Snider GS, 185. Mozumder A. 1974. J. Chem. Phys.
Williams RS. 1998. Science 280:1716–21 60:4300–10
166. Zanardi P, Rossi F. 1998. Phys. Rev. Lett. 186. Noolandi J, Hong KM. 1979. J. Chem.
81:4752–55 Phys. 70:3230–36
167. Welser JJ. Tiwari S, Rishton S, Lee KY, 187. Wang, Y. 1995. Photophysical and pho-
Lee Y. 1997. IEEE Elect. Dev. Lett. tochemical processes in semiconductor
18:278–80 nanoclusters. In Advances in Photochem-
168. Rimberg AJ, Ho TR, Clarke J. 1995. Phys. istry, ed. DC Neckers, DH Volman, G von
Rev. Lett. 74:4714–17 Bünau, 19:179–234. New York: Wiley &
169. Middleton AA, Wingreen NS. 1993. Sons
Phys. Rev. Lett. 71:3198–201 188. Greenham NC, Peng X, Alivisatos AP.
170. Andres RP, Bielefeld JD, Henderson JI, 1996. Phys. Rev. B 54:17628–37
Janes DB, Kolagunta VR, et al. 1996. Sci- 189. Artemyev MV, Bibik AI, Gurinovich LI,
ence 273:1690–93 Gaponenko SV, Woggon U. 1999. Phys.
171. deJongh LJ, ed. 1994. Physics and Chem- Rev. B 60:1504–6
istry of Metal Cluster Compounds. pp. 1– 190. Markovich G, Collier CP, Henrichs SE,
37. Boston: Kluwer Remacle F, Levine RD, Heath JR. 1999.
172. Bach U, Lupo D, Comte P, Moser JE, Ann. Chem. Res. 32:415–23
Weissörtel F, et al. 1998. Nature 395:583– 191. Collier CP, Saykally RJ, Shiang JJ,
45 Henrichs SE, Heath JR. 1997. Science
173. O’Reagan B, Grätzel M. 1991. Nature 277:1978–81
353:737–39 192. Remacle F, Collier CP, Heath JR, Levine
174. Cao F, Oskam G, Meyer GJ, Searson PC. RD. 1998. Chem. Phys. Lett. 291:453–58
1996. J. Phys. Chem. 100:17021–27 193. Shiang JJ, Heath JR, Collier CP, Saykally
175. Colvin VL, Schlamp MC, Alivisatos AP. RJ. 1998. J. Phys. Chem. B 102:3425–30
1994. Nature 370:354–57 194. Markovich G, Collier CP, Heath JR. 1998.
176. Dabbousi BO, Bawendi MG, Onitsuka Phys. Rev. Lett. 80:3807–10
O, Rubner MF. 1995. Appl. Phys. Lett. 195. Medeiros-Ribeiro G, Ohlberg DAA,
66:1316–18 Williams RS, Heath JR. 1999. Phys. Rev.
177. Artemyev MV, Sperling V, Woggon U. B 59:1633–36
1997. J. Appl. Phys. 81:6975–77 196. Weitz I, Sample J, Ries R, Spain E, Heath
178. Schlamp MC, Peng X, Alivisatos AP. JR. 2000. J. Phys. Chem. B. Submitted
Annual Review of Materials Science
Volume 30, 2000

CONTENTS
The Theory of Real Materials, Marvin L. Cohen 1
Tribochemical Polishing, Viktor A. Muratov, Traugott E. Fischer 27
High-Tc Superconductivity in Electron-Doped Layered Nitrides, Shoji
53
Yamanaka
Holographic Polymer-Dispersed Liquid Crystals (H-PDLCs), T. J.
83
Bunning, L. V. Natarajan, V. P. Tondiglia, R. L. Sutherland
Optical Generation and Characterization of Acoustic Waves in Thin
Films: Fundamentals and Applications, John A. Rogers, Alex A. Maznev, 117
Matthew J. Banet, Keith A. Nelson
Structure Evolution During Processing of Polycrystalline Films, C. V.
159
Thompson
Mechanical Behavior of Metallic Foams, L. J. Gibson 191
Annu. Rev. Mater. Sci. 2000.30:545-610. Downloaded from www.annualreviews.org
Access provided by University of Pennsylvania on 04/13/16. For personal use only.

Copper Metallization for High-Performance Silicon Technology, R.


229
Rosenberg, D. C. Edelstein, C.-K. Hu, K. P. Rodbell
The Properties of Ferroelectric Films at Small Dimensions, T. M. Shaw,
263
S. Trolier-McKinstry, P. C. McIntyre
IC-Compatible Polysilicon Surface Micromachining, J. J. Sniegowski, M.
299
P. de Boer
SiGe Technology: Heteroepitaxy and High-Speed Microelectronics, P. M.
335
Mooney, J. O. Chu
Ultrathin Diffusion Barriers/Liners for Gigascale Copper Metallization,
363
A. E. Kaloyeros, E. Eisenbraun
Magnetocaloric Materials, K. A. Gschneidner Jr., V. K. Pecharsky 387
Advances in In Situ Ultra-High Vacuum Electron Microscopy: Growth of
431
SiGe on Si, Ruud M. Tromp, Frances M. Ross
Layered Magnetic Manganites, T. Kimura, Y. Tokura 451
The Electronic Structure of Semiconductor Nanocrystals, Al. L. Efros, M.
475
Rosen
Mechanisms for Enhanced Formation of the C45 Phase of Titanium
Silicide Ultra-Large-Scale Integration Contacts, J. M. E. Harper, C. 523
Cabral Jr.,, C. Lavoie
Synthesis and Characterization of Monodisperse Nanocrystals and Close-
Packed Nanocrystal Assemblies, C. B. Murray, C. R. Kagan, M. G. 545
Bawendi
Extremely High Density Longitudinal Magnetic Recording Media, Dieter
611
Weller, Mary F. Doerner
Low Dielectric Constant Materials for ULSI Interconnects, Michael
Morgen, E. Todd Ryan, Jie-Hua Zhao, Chuan Hu, Taiheui Cho, Paul S. 645
Ho
Device Innovation and Material Challenges at the Limits of CMOS
681
Technology, P. M. Solomon
View publication stats

You might also like