Sie sind auf Seite 1von 49

This article was downloaded by: [Dalhousie University]

On: 20 August 2013, At: 06:38


Publisher: Taylor & Francis
Informa Ltd Registered in England and Wales Registered Number: 1072954 Registered office: Mortimer
House, 37-41 Mortimer Street, London W1T 3JH, UK

Journal of Environmental Science and Health, Part


A: Toxic/Hazardous Substances and Environmental
Engineering
Publication details, including instructions for authors and subscription information:
http://www.tandfonline.com/loi/lesa20

On the Characterization of Environmental


Nanoparticles
a a a
David J. Burleson , Michelle D. Driessen & R. Lee Penn
a
Department of Chemistry, University of Minnesota, Minneapolis, MN, USA
Published online: 28 Mar 2012.

To cite this article: David J. Burleson , Michelle D. Driessen & R. Lee Penn (2004) On the Characterization of
Environmental Nanoparticles, Journal of Environmental Science and Health, Part A: Toxic/Hazardous Substances and
Environmental Engineering, 39:10, 2707-2753, DOI: 10.1081/ESE-200027029

To link to this article: http://dx.doi.org/10.1081/ESE-200027029

PLEASE SCROLL DOWN FOR ARTICLE

Taylor & Francis makes every effort to ensure the accuracy of all the information (the “Content”) contained
in the publications on our platform. However, Taylor & Francis, our agents, and our licensors make no
representations or warranties whatsoever as to the accuracy, completeness, or suitability for any purpose of
the Content. Any opinions and views expressed in this publication are the opinions and views of the authors,
and are not the views of or endorsed by Taylor & Francis. The accuracy of the Content should not be relied
upon and should be independently verified with primary sources of information. Taylor and Francis shall
not be liable for any losses, actions, claims, proceedings, demands, costs, expenses, damages, and other
liabilities whatsoever or howsoever caused arising directly or indirectly in connection with, in relation to or
arising out of the use of the Content.
This article may be used for research, teaching, and private study purposes. Any substantial or systematic
reproduction, redistribution, reselling, loan, sub-licensing, systematic supply, or distribution in any
form to anyone is expressly forbidden. Terms & Conditions of access and use can be found at http://
www.tandfonline.com/page/terms-and-conditions
JOURNAL OF ENVIRONMENTAL SCIENCE AND HEALTH
Part A—Toxic/Hazardous Substances & Environmental Engineering
Vol. A39, No. 10, pp. 2707–2753, 2004

On the Characterization of
Environmental Nanoparticles
Downloaded by [Dalhousie University] at 06:38 20 August 2013

David J. Burleson, Michelle D. Driessen, and R. Lee Penn*

Department of Chemistry, University of Minnesota, Minneapolis, MN, USA

ABSTRACT

The presence and release of nanoparticles into the environment has important
implications for human health and the environment. This article highlights and
describes techniques that are effective in the characterization of anthropogenic
and naturally occurring nanoparticles. Particle attributes like size, size distribu-
tion, shape, structure, microstructure, composition, and homogeneity are
critically important to determining the potential impact of such materials on
health and the environment. Many techniques yield data for a collection of
nanoparticles; while others yield data for individual nanoparticles; and still others
yield data showing the size, distribution of chemical species, and variations in
structure and microstructure for a single nanoparticle. All are important in the
context of environmental nanoparticles. Many of these techniques are comple-
mentary, and depending on the information required, the ideal characterization
usually employs multiple techniques.

Key Words: Nanoparticles; Nanomaterials; Materials characterization; TEM;


Particle size.

*Correspondence: R. Lee Penn, Department of Chemistry, University of Minnesota,


Minneapolis, MN 55455, USA; E-mail: penn@chem.umn.edu.

2707

DOI: 10.1081/LESA-200027029 1093-4529 (Print); 1532-4117 (Online)


Copyright & 2004 by Marcel Dekker, Inc. www.dekker.com
ORDER REPRINTS

2708 Burleson, Driessen, and Penn

INTRODUCTION

Many common products and processes use and produce solid-state nanomater-
ials. The presence and release of such materials into the environment has important
implications for human health (Anastasio[1] and Cass et al.[2] and references
contained therein) and the environment. For example, common products containing
nanoparticulate components include cosmetics, sunscreens, pigments, catalysts,
pharmaceuticals, building materials, and electronic devices. Furthermore, much of
the reactive surface area on and near the Earth’s surface is comprised of
nanoparticles that are the products of physical and chemical weathering of rocks
and minerals, combustion, volcanic eruptions, biomineralization, and precipitation
reactions (e.g., Anastasio[1] and Banfield and Zhang[3]). While such nanoparticles
Downloaded by [Dalhousie University] at 06:38 20 August 2013

represent only a tiny fraction of the Earth’s total mass, they represent a large fraction
of the surface area, which includes solid–solid, solid–liquid, and solid–gas interfaces.
The transport of anthropogenic and naturally occurring nanomaterials in the
environment and in living systems and how such materials facilitate or inhibit the
transport of chemical species is an understudied area that is full of opportunity. A
distinction must be made between anthropogenic nanomaterials, which are
manufactured materials, and naturally occurring nanomaterials. However, deter-
mining whether the source of a particular sample of nanoparticles was anthro-
pogenic or natural, or even biogenic or abiogenic, is often not straightforward.
A great deal of research over the last decade has been dedicated to studying
the health effects of airborne ultrafine particles. Many of these studies deal with
the size fraction commonly referred to as PM10, which usually refers to a mixture
of particulates (minerals, soot, biological particles, sea salt, and other solid-
state particles like ice and solid sulfuric acid) with an average aerodynamic size under
ten micrometers. Ultrafine usually refers to particulates with an aerodynamic size
under 100 nanometers. A major question in the field is whether the impact of
exposure to particles is size dependent, and several publications[4–7] specifically
suggest that health effects due to nanoparticle exposure are indeed size dependent.
For example, Brown et al. showed that lung inflammation resulting from the
inhalation of polystyrene nanoparticles was greatest for the smallest nanoparticles
used.[5]
In addressing the importance of nanoparticles in the environment, it is
important to understand the environment in which the nanoparticles are found or
used. For example, iron metal nanoparticles are finding use for the remediation of
sites contaminated with chlorinated hydrocarbons and other anthropogenic
contaminants.[8–10] In studying the mechanisms by which the nanoparticles are
involved in the chemical transformations of contaminants, it is important to examine
not only the chemical species before, during, and after reaction, but also the
nanoparticles before, during, and after reaction. For example, the reductive
transformation of anthropogenic contaminants like chlorinated hydrocarbons and
nitroaromatic compounds is accompanied by the oxidation of either dissolved
chemical species or solids (e.g., iron metal particles or naturally occurring materials).
Thus, to fully understand the mechanisms of such transformations, both the
molecular and solid-state products must be characterized. The focus of this article is
the characterization of solid nanoparticles.
ORDER REPRINTS

Characterization of Environmental Nanoparticles 2709

Chemistry and Structure of Nanoparticles

The chemical composition, structure, size, and shape of nanoparticles play a


major role in determining their impact on health and the environment. While the
elemental composition of a material or particle is important, there are many cases
where the potential health and environmental effects depend on the oxidation state
of the atoms within the material, the spatial distribution of the elements, and the size
of the particles. For example, Cr(VI) is a known human carcinogen while Cr(III) is a
trace nutrient involved in glucose metabolism.[11,12]
The optical, electronic, magnetic, and physical properties of nanoparticles often
deviate substantially from their bulk counterparts as a result of their small size.[13–17]
For example, the melting temperature of crystalline gold nanoparticles is signifi-
Downloaded by [Dalhousie University] at 06:38 20 August 2013

cantly lower than the melting temperature of bulk, crystalline gold.[18] For a nano-
particle that is five nanometers in diameter, nearly one half of the particle’s atoms
are located at or near the surface. Chemically, surface atoms and ions are distinct
from bulk atoms because they are either coordinatively undersaturated (e.g., in a
vacuum) or bonded to species from the surroundings (e.g., solution species, atoms
from adjacent solid material, or gas molecules) rather than bonded to the next layer
of bulk atoms. Furthermore, the atoms and ions near the surface layer are chemically
distinct from bulk atoms and ions because some of their nearest and next-to-nearest
neighbors are surface atoms.
Attributes such as particle size and shape impact how easily particles are
transported into and throughout the environment and living tissue. The vast
majority of reactions involving nanoparticles occur at the interface between the
nanoparticle and its environment (liquid, solid, or gas). Thus, surface area per unit
mass, porosity, and the condition of the surface (e.g., presence of adsorbates or
surface-bound water) are also significant. In addition, there is a growing body of
evidence that smaller particles are more reactive than larger particles, even when
rates of reaction are normalized to the total amount of surface area present.[19–21]
Furthermore, the crystal structure (e.g., anatase and rutile are two polymorphs of
titanium dioxide), crystallinity, and microstructure (e.g., type and concentration
of dislocations and stacking faults) can affect chemical reactivity. An example of
microstructure-controlled reactivity includes the formation of dissolution etch-pits
originating at dislocations contained at the interface between albite and orthoclase
aluminosilicates, which are two common minerals.[22] Thus, the characterization of
nanoparticle composition, surface area, shape, size, structure, and microstructure as
well as determining the distribution of species throughout a nanoparticle (including
the surface) are critical to evaluating the health and environmental impacts of
naturally occurring and anthropogenic nanoparticles.

General Sample Preparation

Sample preparation is a critical step in the characterization of nanoparticles. A


careful record of sample collection, storage procedures, and preparation steps is
important in order to track potential artifacts. Furthermore, nanoparticles can and
do change structure and composition in response to their environment. For example,
ORDER REPRINTS

2710 Burleson, Driessen, and Penn

placing a hydrated particle into a high vacuum system can dramatically alter both
the composition and structure of the particle. Furthermore, recent work has shown
that the crystal structure of nanoparticles can change as a result of changing the
solvent to which they are exposed.[23] An additional concern is how the particles
change with time. For example, phase transformations and particle growth of
ferrihydrite (a semi-crystalline iron oxyhydroxide mineral) nanoparticles stored in
aqueous suspension is observed even at 2 C.[24]
The quantity of material required can vary from femtograms to grams,
depending on the analysis methods performed and the information desired.
Unfortunately, many nanoparticle characterization methods are destructive, and
the application of some ‘‘nondestructive’’ techniques results in subtle changes in the
sample that can significantly impact results of subsequent analyses. Thus, holistic
Downloaded by [Dalhousie University] at 06:38 20 August 2013

characterization of nanoparticles requires careful planning to ensure the most


accurate results. Lastly, sample collection methods can result in sampling that is not
representative of how the particles occur in the field or are distributed in situ.[25]
Specific sample considerations and preparation procedures are discussed in further
detail in the context of each characterization method presented.
Finally, it is important to evaluate the data produced in the context of the
environment from which the nanoparticles were collected. For example, electron
microscopy performed on iron oxide coated sands collected from an aquifer at
Oyster Bay, VA, shows that the sand coatings are composed of agglomerates of iron
oxide nanoparticles that are attached to Al- and Si-rich nanomaterials, which are
attached to the sand grains. At the centimeter length-scale, the color of the aquifer
sands are strongly banded, reflecting the fact that the concentration of iron oxide
nanoparticles varies as a function of depth.[26] Thus, field data, visual inspection of
the sands, and microscopic characterization yields a holistic description of the
nanoparticles and how they occur in the field.
The organization of this article flows from techniques capable of performing
analyses at the subnanometer length-scale to those methods that produce a single
data point for a sample of nanoparticles. The article is not meant to be an exhaustive
review of nanoparticles in the environment. Rather, it is meant to highlight and
describe techniques effective for characterization of nanoparticles. Table 1 presents
an overview of the techniques presented in this article. Examples from the recent
literature are included where appropriate. Finally, the paper closes with a discussion
of particle size analysis and methods that yield chemical composition data of
individual nanoparticles. It is important to remember that many of these techniques
are complementary, and depending on the information desired, the ideal
characterization commonly employs multiple techniques.

CHARACTERIZATION METHODS: NANOMETER


LENGTH-SCALE

This section presents a discussion of the highest-resolution microscopy


techniques: transmission electron microscopy (TEM) and scanning probe micro-
scopy (SPM). Each technique has advantages and disadvantages with respect to
the characterization of nanoparticles. In both of these high-resolution techniques,
ORDER REPRINTS

Characterization of Environmental Nanoparticles 2711

direct images of individual nanoparticles can be obtained. However, an extremely


small amount of material is characterized, which means that it can be extremely
difficult to ensure that a representative sample is examined. Thus, it is important to
consider the context in which the nanoparticles are important (e.g., the field site or
proximity to a particular nanoparticle source). Finally, sample preparation and
collection and the measurement conditions can often change the nanoparticles or
enrich the sample in certain components of a heterogeneous sample.

Transmission Electron Microscopy (TEM)

Transmission electron microscopy is indispensable when it comes to the


Downloaded by [Dalhousie University] at 06:38 20 August 2013

characterization of nanoparticles and other nano-structured materials. Broadly,


TEM methods can be broken down into three basic areas that strongly complement
one another: (1) imaging at a range of resolutions, where high-resolution images can
directly image the atomic structure of a material; (2) chemical analysis at a spatial
resolution as low as the sub-nanometer range; and (3) electron diffraction, which
probes crystal structure characteristics. For detailed discussions of the construction,
theory, and operation of TEM, the reader is referred to Williams and Carter,[27]
Buseck et al.,[28] Fultz and Howe,[29] and Wang.[30]
Typically, the TEM consists of an electron source (tungsten filament, LaB6
filament or a field-emission gun (FEG)), a high voltage tank (100–1000 kV), a
sample stage, a series of electromagnetic lenses (condenser, objective, and projector
lens systems), a data acquisition system (charge–coupled device (CCD), film, or
video camera), and an analytical system (X-ray spectroscopy or electron energy loss
spectroscopy). A typical TEM operates at a pressure of 109 Torr, which means
that this is inherently an ex situ technique.
When a material is irradiated with an energetic beam of electrons, several types
of primary and secondary signals are produced. Characteristic X-rays (secondary) as
well as elastically and inelastically scattered electrons (primary) are the most
commonly monitored signals in TEM and are the signals discussed here. Analytical
electron microscopy (AEM) refers to the combination of imaging and spectroscopy.
Two types of spectroscopy are commonly employed: X-ray spectroscopy (commonly
referred to as EDS, energy-dispersive spectroscopy; XEDS, X-ray energy-dispersive
spectroscopy; or EDX, energy-dispersive X-ray spectroscopy) and electron energy
loss spectroscopy (EELS).

Imaging

The major advantage of using TEM to characterize nanoparticles is that a direct


image of the nanoparticle is obtained. The interaction of the electron beam with the
specimen produces both amplitude and phase variations in the electron wave that are
governed by quantum mechanical diffraction theory. Three types of contrast are
observed in TEM images. Mass thickness contrast is sensitive to both the thickness
of the sample (i.e., the amount of material through which the electron beam traveled)
and the atomic number of the atoms (heavier atoms scatter electrons more strongly
than lighter atoms). Diffraction and phase contrast are a direct consequence of the
Downloaded by [Dalhousie University] at 06:38 20 August 2013

Table 1. Techniques used for nanoparticle characterization, organized according to their appearance in this article.
2712

Spatial resolution,
Main minimum particle size, Sample
Acronym(s) Meaning informationa or detection limit requirementsb Notes:

Nanometer length-scale
STEM Scanning Transmission I, S, D, P, APS <0.1 nm spatial resolution Sub-mg, a In combination with EELS/EDS: C,
Electron Microscopy Ox, and CB is possible. Can
quantify the size of an individual
particle
TEM Transmission Electron I, S, D, P, APS >0.1 nm spatial resolution Sub-mg, b In combination with EELS/EDS: C,
Microscopy Ox, and CB is possible. Can
quantify the size of an individual
ORDER

particle
HRTEM High-Resolution TEM I, S, D, P, APS 0.1 nm spatial resolution Sub-mg, b In combination with EELS/EDS: C,
Ox, and CB is possible. Can
quantify the size of an individual
particle
REPRINTS

SAED Selected Area Electron D, P, S 0.1 mm diameter area Sub-mg, b Scattering volume is selected by use of
Diffraction a selected area aperture
CBED Convergent Beam Electron D, P, S 1 nm diameter area Sub-mg, a Scattering volume is selected by
Diffraction focusing the electron beam
EELS Electron Energy Loss C, Ox, CB  0.5 nm spatial resolution Sub-mg, b Smallest volume analyzable is
Spectroscopy determined by the size of the
electron probe. Single atom
detection possible with some
elements
EDS Energy Dispersive C  0.5 nm spatial resolution Sub-mg, b Smallest volume analyzable is slightly
Spectroscopy Also larger than the projected volume
EDX, XEDS defined by the size of the electron
probe
Burleson, Driessen, and Penn
Downloaded by [Dalhousie University] at 06:38 20 August 2013

EFTEM Energy-Filtered TEM I, S, D, P, C, Ox, CB 0.5 nm for C, Ox, and CB Sub-mg, b Energy-filtered images are formed
using electrons that have lost a
specified amount of energy by
coupling EELS with TEM
HAADF High-Angle Annular Dark I, S, D, C 0.1 nm Sub-mg, a Also known as Z-contrast imaging.
Field Single heavy atom imaging is
possible using HAADF
AEM Analytical Electron I, S, D, P, C, Ox, CB, APS 0.5 nm spatial resolution Sub-mg, b AEM refers to the combination of
Microscopy for analyses, 0.1 nm for imaging nd analytical techniques
imaging (e.g., EELS and EDS)
ETEM Environmental TEM I, S, D, P, C, Ox, CB >0.1 nm spatial resolution Sub-mg, stable under Semi- in situ measurements possible
ORDER

electron beam. (sample chamber at 1–10 Torr


pressure)
SPM Scanning Probe Microscopy I, APS, other 0.1 nm spatial resolution Sub-mg, c Can quantify the size of an individual
particle. Can also characterize
Characterization of Environmental Nanoparticles

electrical and mechanical properties


REPRINTS

and measure forces of interaction


between sample and tip
AFM Atomic Force Microscopy I, APS, other 0.1 nm spatial resolution Sub-mg, c Can quantify the size of an individual
particle. Can also characterize
electrical and mechanical properties
and measure forces of interaction
between sample and tip
NSOM Near-field Scanning Optical I, CB 30 nm spatial resolution Thin samples Often used in combination/
Microscopy (<200 nm) concurrently with AFM
STM Scanning Tunneling I 0.1 nm spatial resolution Sub-mg, adhered to Not common for the characterization
Microscopy substrate, conduc- of environmentally relevant
tive only. nanoparticles

(continued)
2713
Downloaded by [Dalhousie University] at 06:38 20 August 2013

Table 1. Continued.
2714

Spatial resolution,
Main minimum particle size, Sample
Acronym(s) Meaning informationa or detection limit requirementsb Notes:

Micrometer length-scale
SEM Scanning Electron I, APS 1 nm–1 mm spatial resol- Sub-mg, semicon- Can quantify the size of larger
Microscopy ution depending on signal ductive or con- individual particles. In combination
monitored ductive with Auger Spectroscopy: C, Ox,
and CB. In combination with EDS:
C. In combination with EBSD: S, P
SE Secondary Electrons I, APS 1 nm spatial resolution SE are commonly detected in SEM.
SE intensity maps yield mainly
ORDER

topographical information
BSE Back-scattered Electrons I, some C 1 mm spatial resolution BSE are commonly detected in SEM.
BSE intensity maps yield mainly
atomic number (Z) contrast
EBSD Electron Back-scatter S, P, crystal 1 mm3 Used in combination with SEM: S, P
REPRINTS

Diffraction orientation
AES Auger Electron C, Ox, CB  1–2 nm spatial resolution Sub-mg, semicon- Extremely surface sensitive technique.
Spectroscopy possible. ductive or con- Often used in combination with
ductive SEM. Mostly used for C
ESEM Environmental SEM I, APS 3–4 nm spatial resolution Same as above Semi- in situ measurements possible
(sample chamber at 1–10 Torr
pressure)
PIXE Particle Induced X-ray C 1 mm3 Sub-mg Trace element detection limits on the
Emission order of 1015 g In situ
measurements possible
XPS X-ray Photoelectron C, Ox, CB 1 mm spatial resolution Sub-mg Extremely Surface Sensitive Technique
Spectroscopy possible
Burleson, Driessen, and Penn
Downloaded by [Dalhousie University] at 06:38 20 August 2013

Bulk analyses (These methods produce an average result for a sample of nanoparticles)
AA Atomic Absorption Also C 10–100 ppb detection limit Sub-mg Average compositional analysis.
AAS (see text) Sample prep. generally requires
sample dissolution
AES Atomic Emission Spectros- C 10–100 ppb detection limit Sub-mg Average compositional analysis.
copy Sample prep. generally requires
sample dissolution
ICP-AES Inductively Coupled Plasma C 1–100 ppb detection limit Sub-mg Average compositional analysis.
AES Sample prep. generally requires
sample dissolution
ICP-MS ICP Mass Spectrometry C 1–10 ppt detection limit micrograms Average compositional analysis.
Sample prep. generally requires
sample dissolution
XRD X-ray Diffraction S, D, P, APS 1–3 wt % detection limit Milligrams Ideal for crystalline nanoparticles.
ORDER

Can be operated in micro-


diffraction mode (10 mm spatial
resolution)
XAS X-ray Absorption Spectros- C, Ox, CB, S 10–100 ppm detection 5–10 mg Ideal for average C, Ox, and CB
Characterization of Environmental Nanoparticles

copy limit characterization of nanoparticle


REPRINTS

sample
EXAFS Extended X-ray Absorption S, CB 5–10 mg XAS method: ideal for probing the
Fine structure arrangement, number, and identity
of nearest neighbor atoms around
the absorbing element
XANES X-ray Absorption Near Edge C, Ox, CB Milligrams XAS method: sensitive to the
Structure oxidation state and symmetry
around the absorbing element
SEXAFS Surface Extended X-ray C, Ox, CB, S Thin film Surface sensitive (probes outer 1 nm
Absorption Fine Structure of material)
IR, FTIR Infrared Spectroscopy; P, CB Milligrams Vibrational spectroscopy.
Fourier Transform IR Incompatible with wet systems
Raman Raman Spectroscopy P, CB Milligrams Vibrational spectroscopy. Compatible
with aqueous suspension and wet
nanoparticle samples
2715

(continued)
Downloaded by [Dalhousie University] at 06:38 20 August 2013

Table 1. Continued.
2716

Spatial resolution,
Main minimum particle size, Sample
Acronym(s) Meaning informationa or detection limit requirementsb Notes:

Magnetic Mössbauer Spectroscopy P, APS, Ox P 10 mg tens of mg These methods are useful for the
properties Magnetometry characterization of iron rich
environmental nanoparticles, even
in extremely heterogeneous samples
BET Brunauer, Emmett, and Total surface area Up to thousands of m2/g 100 mg Method is most sensitive for high
Teller surface area materials (small particle
size). Method can also be used to
characterize porosity of
ORDER

nanoparticles
PCS Photon Correlation APS 3 nm–mm particle size Sub-mg in liquid Average particle size and size
Spectroscopy or DLS suspension distribution obtainable in situ
Single particle analyses
ATOFMS Aerosol Time of Flight Mass APS, C 3 nm–mm particle size <pg Can analyze a single nanoparticle for
REPRINTS

Spectroscopy both size and C


DMA Differential Mobility APS 3 nm–mm particle size <pg Can analyze a single nanoparticle for
Analyzer size
TDMA Tandem DMA APS, other 3 nm–mm particle size < pg Can analyze a single nanoparticle for
size. Tandem systems can involve a
DMA in combination with a wide
array of second techniques
(e.g., ATOFMS)
a
I ¼ image, S ¼ structure (e.g., amorphous or crystalline), D ¼ defect analysis, P ¼ Phase identification, C ¼ chemical compositional analysis,
Ox ¼ oxidation state, CB ¼ chemical bonding environment, and APS ¼ average particle size.
b
a ¼ stable under focused electron probe and in high vacuum, b ¼ stable under electron beam and in high vacuum, c ¼ adhered to substrate,
conductive or nonconductive.
Burleson, Driessen, and Penn
ORDER REPRINTS

Characterization of Environmental Nanoparticles 2717


Downloaded by [Dalhousie University] at 06:38 20 August 2013

Figure 1. TEM images of an ultramicrotomed section of a sulfate-reducing bacteria (SRB)


biofilm from the Piquette Mine, near Tennyson, WI. The mine is a dolomitic stratabound
Pb/Zn deposit that was closed 33 years ago and allowed to flood. Near wood timbers,
localized white biofilms have developed. These are dominated by SRB and concentrate
micron-sized spheroids that are agglomerates of zinc sulfide nanocrystals of ZnS. On the left-
hand side is low magnification image (taken using a JEOL1010 TEM operated at 80 kV) of a
SRB and several spheroids of zinc sulfide. On the right-hand side is a high-resolution TEM
image (taken using a JEOL4010 HRTEM) of the zinc sulfide nanoparticles that comprise the
spheroids. Images are courtesy of J. Moreau and J. F. Banfield at the University of
California—Berkeley.

fact that electrons have wave character. Diffraction contrast depends on the
orientation of a crystal with respect to the electron beam. At certain orientations
(e.g., when the Bragg condition is satisfied, see below), scattering due to diffraction is
very strong, which produces dark regions in the TEM image. Phase contrast is the
consequence of phase shifts in the electron beam as a result of interacting with
the sample (e.g., Fig. 1). The contrast observed in atomic structure images and lattice
fringe images belongs to this category. When atomic structure images are compared
to image simulations, direct information about the distribution of atoms in a sample
can be obtained. Image simulations are produced by modeling imaging conditions
(e.g., focus, stigmation, beam tilt, sample tilt, and aperture selections), the
microscope specifications (e.g., spherical aberration constant), and the structure of
the sample. Through careful comparison of the experimental and simulated images,
which can be performed quantitatively, it is possible to determine which features in
an image correspond to actual positions of atoms and/or clusters of atoms.
Using TEM images, particle size, size distribution, and morphology can be
evaluated. In general, measuring particle sizes using TEM images can be performed
with some accuracy (5%) when good standards are employed.[27] At the highest
ORDER REPRINTS

2718 Burleson, Driessen, and Penn

resolutions, crystalline nanoparticles with known crystal structure provide a


convenient means with which calibration can be performed (i.e., lattice fringe
images can be used as an internal standard).
Because the TEM image is a two-dimensional representation of a three-
dimensional object, the sample must be examined in at least two orientations in
order to obtain a three-dimensional picture of an individual nanoparticle or
aggregate of nanoparticles. If two or more nanoparticles are overlapping or joined in
some fashion, it can be difficult to determine which nanoparticle is on top or whether
particles are joined by chemical bonds or are simply overlapping. If the particle size
is large enough, then SEM (scanning electron microscope, see below) images can
strongly complement the TEM images.
Downloaded by [Dalhousie University] at 06:38 20 August 2013

Diffraction

The electron diffraction pattern originates from the volume of sample traversed
by the electron beam. The wave character of electrons means they can be diffracted.
The electron diffraction pattern is similar to an X-ray diffraction pattern (discussed
below) in that diffraction occurs when the Bragg condition is satisfied. The Bragg
equation (n ¼ 2 d sin ) relates the wavelength () of the electrons, the angle () of
the electron beam with respect to a set of diffracting planes, and the spacing (d)
between those planes, which is determined by the crystal structure of the material
examined. One challenge when working with electron diffraction is that electrons are
often multiply scattered by the sample, which can complicate the interpretation of
electron diffraction patterns.
Using electron diffraction, it is possible to distinguish between amorphous and
crystalline materials and often possible to identify crystalline materials by measuring
d-spacings and examining the symmetry of the pattern. Furthermore, a diffraction
pattern can be obtained from an individual nanoparticle or, in some cases, from
a portion of a nanoparticle. The diffracting volume can be limited by using selected-
area electron diffraction (SAED), which employs an aperture in order to limit the
volume of sample from which the pattern is generated, or convergent beam electron
diffraction (CBED), which involves focusing the electron beam into a narrow beam
(as narrow as a fraction of one nanometer in diameter). A major advantage of
electron diffraction in the TEM is that it is possible to directly image the particle or
portion of a particle from which a diffraction pattern is taken. In addition, by using
imaging and diffraction, the size and size distribution of crystalline nanoparticles
along specific crystallographic directions can be evaluated.

Spectroscopy

TEMs use an energetic beam of electrons, which, as mentioned above, interact


with the specimen to produce a number of primary and secondary signals. When an
incoming electron ionizes an atom through the removal of an inner shell electron, a
characteristic X-ray can be emitted when an electron from a higher energy, outer
shell fills the inner shell vacancy. The energy of the emitted X-ray corresponds to the
ORDER REPRINTS

Characterization of Environmental Nanoparticles 2719

energy difference between the two electron energy levels, and it is characteristic of
the element ionized. Furthermore, the primary electron loses an amount of energy
that is equal to the energy required to remove the inner shell electron.
EDS (or XEDS or EDX) measures the intensity and energy of X-rays generated
from the specimen. The EDS detector is positioned very close to the sample and
typically kept at liquid nitrogen temperature and under vacuum. Thus, a window is
often employed in order to prevent condensation of gaseous species onto the detector
semiconductor. A consequence of using a window is substantially reduced sensitivity
to characteristic X-rays from the lighter elements. The detection limits can be as low
as sub-parts per thousand for elements heavier than sodium and for which very little
EDS peak overlap is observed.[27,31] Quantitative results can be obtained by
generating k-factors, which relate the intensity of the EDS peak to concentration
Downloaded by [Dalhousie University] at 06:38 20 August 2013

through the analysis of well-understood standards. The reader is referred to Joy


et al.[32] and Garratt-Reed and Bell[33] for a more detailed treatment of the theory
and application of EDS in the electron microscope.
Electron energy loss spectroscopy measures the energies of electrons after
interacting with the specimen. The energies of these electrons are classified according
to the energy lost. For example, the electron that resulted in the ionization of a
particular atom will have lost an amount of energy equal to the energy required to
remove that inner shell electron. Thus, similar compositional information can be
obtained as in EDS. However, the greatly increased energy resolution of EELS
(fraction of an eV) vs. EDS (>100 eV) means that information about bonding,
electronic structure (e.g., oxidation state), and average bond distances can be
obtained using EELS. An additional advantage of EELS over EDS is that beam
broadening is not a major concern when using EELS. When a beam of electrons is
transmitted through a specimen, beam spreading occurs as a result of scattering
events. Thus, the characteristic X-rays generated come from a volume of material
that is larger than the projected volume without beam broadening. In EELS, the use
of an entrance aperture excludes those electrons that have been elastically scattered
through large angles. In other words, the spatial resolution achievable with EELS is
superior to that achievable with EDS. On the other hand, a major advantage of EDS
over EELS is the relative ease with which the technique can be applied.
Theoretically, EELS can be used to detect all elements. In practice, however,
working with hydrogen and helium can be challenging because other features in the
spectrum often overlap with features produced by electron interactions with these
two elements. The detection limit for EELS is approximately 1021 g of material,[31]
which corresponds to as little as one heavy element atom. Quantitative composi-
tional analyses are possible with the use of appropriate standards (e.g., thickness and
chemical bonding of the element of interest must match for sample and standard),
and the accuracy is 1–2%.[31] For a detailed treatment of the theory and operation of
EELS, the reader is referred to texts by Egerton[34] and Ahn et al.[35]

Advanced Techniques

Many TEMs can be operated in STEM (scanning transmission electron


microscopy) mode, which rasters a focused beam across the sample. Dedicated
ORDER REPRINTS

2720 Burleson, Driessen, and Penn


Downloaded by [Dalhousie University] at 06:38 20 August 2013

Figure 2. High-resolution zero-loss TEM images of coating material prepared using a gentle
abrasion method to remove ‘‘iron oxide’’ coatings from sand grains from Oyster Bay Virginia.
The white arrow in the left-hand image serves to highlight an agglomerate of goethite
(a-FeOOH) nanoparticles. EDX and electron diffraction results demonstrate that the
agglomerates are composed of goethite nanoparticles. The middle image shows the
agglomerate at higher magnification. The right-hand image is a micrograph of a different
agglomerate in which a significant degree of preferred orientation of the primary goethite
nanoparticles is observed. The inset in the right-hand image is a fast-fourier transform, which
serves to highlight the preferred orientation of the primary nanoparticles. These images were
taken using a 300 kV FEI CM300 FEG TEM-STEM.

STEMs only operate in this mode, and the achievable image resolution in STEM
mode is superior to the achievable image resolution in TEM mode. STEM can be
combined with spectroscopic methods (EELS or EDS) to produce spectrum images
(e.g., Fig. 2) and HAADF (high-angle annular dark field) to produce ‘‘Z-contrast’’
images. Spectrum images are four-dimensional images that contain, in addition to
contrast information, a spectrum at each pixel. Z-contrast imaging takes advantage
of mass thickness contrast to produce images in which the contrast is directly related
to the atomic weight of the atoms scattering the electron beam. Images of individual
heavy atoms on low atomic weight materials can be produced. Electron holography
is a specialized TEM technique that can produce images with enhanced
resolution.[27,36,37] Lorentz microscopy is yet another specialized TEM or STEM
technique that enables the user to directly image the magnetic microstructure of the
sample.[27,38]

Sample Preparation

In general, nanoparticles are deposited onto a support film (such as a polymer or


amorphous carbon film) that is supported by a copper (or other metal) grid. For
TEM, the sample must be conductive. If the nanoparticles themselves are not
conductive, then coating the sample with amorphous carbon or using an amorphous
carbon support film provides the conductivity. Unfortunately, the image will then
include both the film and the nanoparticle. Often, the microscopist needs to collect
high-resolution images of just the nanoparticle as opposed to the nanoparticle and
ORDER REPRINTS

Characterization of Environmental Nanoparticles 2721

the support film. One alternative is to use holey or lacey films, which have many
small holes, in the hope that some particles will either adhere to the side of a hole or
partially overlap with a hole.
Particles can be deposited onto the grid by several methods. An easy method
involves suspending particles in a solvent that does not dissolve the support film and
dipping the grid into the suspension or placing a single drop of the suspension onto
the grid (on the support film side) and then allowing solvent to evaporate. Another
method is to place the grid in an aerosol stream. Ideally, the amount of coverage
should be approximately 10% or less, and the ideal suspension concentration or time
in the particle stream can usually be estimated. Still another method is to dip the grid
into a dried powder of nanoparticles and shake the excess off. The preparation
method of choice will depend on the sample and the information desired. It is
Downloaded by [Dalhousie University] at 06:38 20 August 2013

important to pay close attention to the sample preparation method used since many
methods can introduce substantial artifacts (e.g., aggregate fragmentation upon
impact with support film or precipitation of salts and nanoparticle aggregation upon
drying a droplet of suspension on support film). Lastly, the ideal TEM sample is less
than 200 nm thick, which is usually not a concern with nanoparticle samples.
However, in the event that the particles are fairly large, they can be suspended in an
epoxy, and mechanical polishing or ion milling of the resulting particle mount can be
used to prepare an electron transparent sample. Additional and more complex
methods of sample preparation are available (e.g., ion millers and dimplers), and the
reader is referred to Williams and Carter for additional details.[27]
Sample preparation techniques to avoid or minimize aggregation include the
addition of a volatile acid or base (e.g., acetic acid or ammonia) to an aqueous
suspension prior to dipping or drop placement. If a volatile acid or base is used, then
the acid or base will evaporate along with the water. Care must be taken that the acid
or base selection is appropriate for the nanoparticles to be examined (e.g., solubility
considerations). This can reduce aggregation if the pH of the solution can be moved
significantly away (at least two or three pH units) from the average isoelectric point
of the nanoparticles (for a discussion of isoelectric points, the reader is referred to
Hiemenz and Rajagopalan.[39]). In addition, if one must prepare a suspension from a
dry powder, sonication can be used to help break up aggregates.
If aggregation of particles is observed in TEM images, it does not necessarily
imply that the particles were strongly aggregated in situ. Also, if particle aggregates
are fragmented on the TEM grid, it does not necessarily imply that the aggregates
are fragmented in situ. For example, aerosol soot particles that were deposited from
an aerosol stream onto a TEM grid fragmented upon collision with the support film
(Fig. 3 and Park[40]). Thus the average aggregate size measured using TEM was
substantially smaller than the in situ aggregate size.
One of the major disadvantages of TEM for the characterization of
nanoparticles is that it is inherently an ex situ method. The sample must be dried
and inserted into a high vacuum, which makes the study of materials sensitive to
vacuum conditions (e.g., hydrated materials) challenging. Both sample preparation
and insertion into the TEM can introduce artifacts that must be considered. In fact,
Zhang et al. recently demonstrated that the crystal structure of zinc sulfide nano-
particles changes as a result of changing the nature of the molecules surrounding the
particle.[23] They showed that the binding of water molecules to zinc sulfide
ORDER REPRINTS

2722 Burleson, Driessen, and Penn


Downloaded by [Dalhousie University] at 06:38 20 August 2013

Figure 3. TEM images of mobility-classified soot nanoparticle (220 nm mobility size, John
Deere engine, 50% load, 1400 rpm, EPA fuel (360 ppm S)). White arrowheads in the left-hand
image serve to highlight ‘‘fragments’’ that were produced as a result of particle impact onto the
carbon film support. The right-hand image is a higher magnification image of two such
fragments. This image was taken using a JEOL 1210 TEM and at a slight underfocus in order
to highlight particle edges. Images are courtesy of Kihong Park.

nanoparticles results in a structural transformation. An environmental transmission


electron microscope (ETEM) allows the user to characterize a sample in an
environment that is isolated from the vacuum. In an ETEM, samples can be imaged
in controlled atmospheres (up to several Torr total pressure). Recent work has
characterized structural transformations in nanoparticles under controlled environ-
ments[41] and even while reacting with gases.[42] A major challenge of working with
an environmental cell is the greatly increased inelastic scattering from the gas atoms
and molecules. However, the use of an energy-filter can eliminate these inelastically
scattered electrons from electron diffraction patterns and images.
A final and important consideration when using TEM to image nanoparticles is
that only a small fraction of nanoparticle sample can be examined. Furthermore,
some sample preparation methods may enrich the sample in one type of
nanoparticle. Thus, the user must take great care to ensure that a representative
sample is examined. If, for example, 99.9% (by number) of a particular sample of
nanoparticles have diameters of 5–10 nm and 0.1% of the particles have diameters of
100–200 nm, then for every one thousand particles examined, the probability is that
only one large particle will have been examined. However, from the perspective of
mass, the majority of the sample is comprised of the larger particles. If a
microscopist examines only one hundred particles, the resulting description of the
sample could be inaccurate. The point here is that TEM has the major advantage of
being the only technique that can image individual nanoparticles at high-resolution,
but it has the major disadvantage of sampling only a relatively small number of
ORDER REPRINTS

Characterization of Environmental Nanoparticles 2723

nanoparticles. In fact, according to the National Bureau of Standards, a minimum of


10,000 images of nanoparticles should be examined in order to ensure statistical
validity.[43]
Using TEM methods, nanoparticles can be characterized in terms of their crystal
structure (or lack thereof) and chemistry (composition and bonding information) at
the nanometer length-scale; and particle size, size distribution, and morphology can
be determined. Major advantages of using TEM methods to characterize nano-
particles include: (1) individual nanoparticles can be directly imaged, (2) chemical
analysis can be performed on individual nanoparticles at a spatial resolution of less
than one nanometer, (3) electron diffraction patterns can be collected from
individual nanoparticles, (4) nanoparticles from which diffraction patterns and/or
spectra are collected can be directly imaged, and (5) a relatively large number of
Downloaded by [Dalhousie University] at 06:38 20 August 2013

nanoparticles can be directly examined. Major disadvantages of using TEM methods


include: (1) images produced are two-dimensional representations of three-
dimensional objects; (2) TEM is inherently an ex situ technique in that the sample
is usually examined under high-vacuum conditions; (3) many samples change or are
damaged readily upon exposure to the electron beam; (4) a tiny fraction of the
sample is examined; and (5) sample preparation can introduce substantial artifacts.

Scanning Probe Microscopy (SPM)

The primary SPM discussed here is atomic force microscopy (AFM) because it is
the most commonly applied SPM for environmental nanoparticles. The AFM
measures the forces of interaction between a sharp tip (radius of curvature
usually 10 nm) and the sample by rastering the tip over the sample surface or
rastering the sample beneath the tip (using piezoelectric tubes). Force measurements,
force maps, and topographic images with vertical and lateral resolutions in the sub-
nanometer range can be obtained. AFM is useful for imaging the size, shape, and
topography of nanoparticles in situ and, with modified AFM tips, measuring the
forces of interaction between nanoparticles as a function of environment.
The AFM tip is attached to a cantilever with an extremely small spring constant,
and the cantilever deflects when the tip is attracted to or repulsed from the sample
surface. The deflection is sensed using an optical lever, which usually consists of a
red laser reflected off the back-side of the cantilever and onto a position sensitive
photodiode. Deflection is monitored as a function of position, and topographic and
force maps can be produced. In contrast to electron beam methods, AFM can be
routinely performed in situ and with both conductive and non-conductive samples.
For detailed discussions of the construction, theory, and operation of an AFM, the
reader is referred to excellent discussions by Maurice,[44] Friedbacher and Fuchs,[45]
and Sarid.[46]
In general, AFMs can be operated in three modes: contact, noncontact tapping,
and contact tapping modes. In contact mode, the tip is essentially dragged across the
surface, and a topographic map is generated. In tapping mode, the cantilever is
oscillated near its first bending mode resonance frequency with an amplitude of
approximately one to 200 nm. In noncontact mode, this oscillating tip is rastered at a
distance of approximately 10 nm above the surface, and the oscillation amplitude is
ORDER REPRINTS

2724 Burleson, Driessen, and Penn

usually 1–10 nm. In contact tapping mode, the amplitude of tip oscillation is
significantly larger, 20–200 nm, and the tip makes contact with the sample at the
bottom of each oscillation. Tip sample interactions result in measurable changes in
the amplitude and phase of the oscillation as well as the resonant frequency of the
oscillation. All can be measured and used to generate force and topographic maps of
a sample surface. By monitoring different modes of tip deflection, properties like
relative adhesion, hardness, and, in some cases, chemical composition can be
characterized. If the interfacial area between tip and surface and the spring constant
of the cantilever are known, quantitative force measurements can also be performed.
Atomic force microscopy is capable of imaging at sub-nanometer spatial
resolution, and, under ideal conditions (e.g., sharp tip, clean surface), it is theoretically
possible to image the atomic structure of a surface. In terms of particle size and shape
Downloaded by [Dalhousie University] at 06:38 20 August 2013

measurements, the shape of the tip must be well known and good standards must be
employed. Tip shape can be determined by imaging a sharp object. If, for example,
the radius of curvature is very large, then the sharp object will appear large in the
AFM image. If, for example, the tip has a double apex, then the sharp object will be
doubly imaged. Furthermore, if a nanoparticle detaches from the substrate and
attaches to the tip, substantial artifacts will be introduced into the AFM force
measurements and topographic maps. This is usually marked by a sudden loss in
resolution in the AFM image or a dramatic change in force measurements. Tapping
mode is particularly useful in imaging nanoparticles because it minimizes the lateral
force between tip and sample, which reduces the chance that particles will be moved by
the tip. Finally, significant damage can occur to samples that are substantially softer
than the tip, especially when the AFM is operated in contact mode. Conversely,
significant tip damage can occur when imaging materials that are substantially
harder than the tip. The use of tapping modes can substantially reduce such damage.
When imaging in air, capillary forces (due to surface sorbed water) can be
substantial. The most effective method of avoiding complications due to capillary
forces is to image using a fluid cell. The environment of the fluid cell can be
controlled (e.g., anaerobic conditions, choice of solvent) and the fluid composition
varied over time (flow through construction). Thus, dynamic changes in surfaces
undergoing reaction can be quantitatively measured. For example, Nagy et al.
examined the kinetics of gibbsite (Al(OH)3) growth on various substrates using
AFM[47]; Sutheimer et al. measured the dissolution rates of well and poorly
crystallized kaolinite particles in aqueous solutions[48]; and Teng et al. examined
calcite growth in various aqueous solutions.[49]
A major drawback of AFM is chemical information cannot be obtained from
AFM images directly. However, relative differences in chemistry can be determined
by using chemically modified AFM tips, and this method is often termed chemical
force microscopy.[50–52] In magnetic force microscopy (MFM), the tip is coated with
a magnetic material, which enables the user to characterize the magnetic domain
structure of a sample. For example, Proksch et al. used MFM to characterize
magnetic nanoparticles in a magnetotactic bacterial cell.[53] Lateral force microscopy
refers to monitoring the twisting motion of the cantilever as the tip is scanned over
the surface, which enables the user to characterize variations in friction between the
sample and the tip. Furthermore, a relatively new tapping technique looks at the
phase shift in the oscillation of the cantilever in comparison to the applied oscillation
ORDER REPRINTS

Characterization of Environmental Nanoparticles 2725

and enables the user to characterize some specific materials properties like visco-
elasticity.[54] Lastly, near-field scanning optical microscopy (NSOM or SNOM) is
a more recent scanning probe technique that uses light channeled through a
small aperture at the end of a quartz AFM tip.[55–58] Transmitted or fluorescent
light can be used to obtain chemical information and image the sample with better
spatial resolutions (down to 1 nm in some cases) than normal (far-field) optical
microscopy.
A related technique is scanning tunneling microscopy (STM), which monitors
the tunneling current between a conductive tip (usually a sharp metal tip) and a
conductive or semiconductive sample as the tip is rastered over the sample. STM is
capable of achieving molecular- to atomic-scale resolution and is sensitive to the
chemical identity of the atoms and molecules at the surface.[45,59]
Downloaded by [Dalhousie University] at 06:38 20 August 2013

Sample Preparation

Nanoparticles must be strongly adhered to a substrate, and care must be taken


to choose adhesives and substrates that are unlikely to introduce artifacts by causing
sample or tip changes that could complicate the AFM measurements. For example,
if nanoparticles are laid down on a thin layer of epoxy and then heat cured, the heat
treatment could substantially change the nanoparticles. Concurrently, the tip must
not be contaminated or altered as a result of interacting with the substrate. For
example, if a thin layer of soft material is used to adhere nanoparticles to a substrate,
the soft material could essentially stick to the AFM tip, causing major image
artifacts. Lastly, when using a fluid cell, the substrate must not participate in the
chemical reactions of interest or change such that the nanoparticles are no longer
strongly adhered. Bickmore et al. discusses the use of AFM to characterize clay
particles, and a substantial portion of the article specifically addresses sample
preparation.[60] Their work addresses the use of electrostatic attraction (e.g., using
fluid conditions that result in positively charged particles and a negatively charged
substrate) and adhesion to a thermoplastic film.
Disadvantages of AFM for the characterization of nanoparticles include the lack
of direct information about chemical composition, tip limitations (including
sharpness and shape), potential sample and tip damage, potential displacement of
nanoparticles, and data interpretation (i.e., quantifying the forces of interaction
between tip and sample). Advantages include the fact that measurements can be
performed in situ and on nonconducting samples, that measurements can be
performed on surfaces undergoing reaction, the sub-nanometer spatial and vertical
resolution, and the inherent surface sensitivity of the method.

CHARACTERIZATION METHODS: MICROMETER


LENGTH-SCALE

This section presents three techniques that can access length-scales intermediate
between the nanometer and bulk scales. The first, SEM (Scanning Electron
ORDER REPRINTS

2726 Burleson, Driessen, and Penn

Microscopy), fits into both the nanometer and micrometer length-scale groups. It is
ideal for imaging particles in the tens of nanometers to micrometers range and, when
equipped with the appropriate detector, can often be used to perform chemical
analyses at a spatial resolution in the sub-micron range. The second, PIXE (Proton
Induced X-ray Emission), is less common but is ideal for performing trace chemical
analyses down to a spatial resolution of approximately one micrometer. The third,
XPS (X-ray Photoelectron Spectroscopy) is also less common but is ideal for
performing chemical analyses and obtaining information about the chemical state
(e.g., oxidation state) of elements in a sample.

Scanning Electron Microscopy (SEM)


Downloaded by [Dalhousie University] at 06:38 20 August 2013

Typically, SEM consists of an electron source, a sample stage, electromagnetic


lenses, scanning coils, detectors, and a data acquisition system. In addition, SEMs
often have EDS systems for the characterization of chemical composition. A typical
SEM operates at pressures less than 105 Torr to minimize the scattering of electrons
by gas molecules and to prevent electron source failure. Several good references and
recent reviews on SEM can be found in the literature.[61–64]
When a material is irradiated with an energetic beam of electrons, several
primary and secondary signals are produced. Characteristic X-rays, backscattered
electrons (BSE), and secondary electrons (SE) are the most commonly monitored
signals in SEM and are the signals discussed here. Contrast maps produced using SE
intensity yield mainly topographical information while contrast maps produced
using BSE intensity yield mainly atomic number (Z) contrast. Finally, contrast maps
produced using specific energies of characteristic X-rays yield information about the
distribution of a particular element in the sample.
As in TEM, elemental analysis by EDS can be performed in the SEM by
monitoring the energy and intensity of X-rays emitted from the sample.[27,61–64]
Beam broadening is the consequence of electron scattering in the sample and results
in reduced spatial resolutions in EDS. For the very thin samples typically used in
TEM, beam broadening is a minor concern. However, SEM samples are usually
larger and thicker than TEM samples and the achievable spatial resolution for
samples that are several microns thick (e.g., thick layer of nanoparticles on a
substrate) or more is typically  500 nm.

Advanced Techniques

Crystallographic and texture information can be obtained by employing electron


backscatter diffraction (EBSD), which is a consequence of the diffraction of
backscattered electrons from within the sample.[65–68] Additional chemical and
structural information can be obtained for some materials by another technique
known as cathodoluminescence (CL).[61,69,70] Electrons from the beam may excite the
sample’s electrons to higher electronic states. When the excited electrons relax back
to a lower energy state, a photon can be emitted and the energy of the emitted
photon is extremely sensitive to composition and the presence of impurities and
ORDER REPRINTS

Characterization of Environmental Nanoparticles 2727

defects. Lastly, Auger electron spectroscopy (AES) can be performed in an SEM


equipped with the appropriate detector. The energy of Auger electrons is
characteristic of the element from which they are emitted and sensitive to oxidation
state and chemical bonding environment. This technique samples the top 2 nm of
material and is ideal for characterizing surfaces,[71] which can be useful for
characterizing the difference between the surface and interior chemistry of larger
nanoparticles (i.e., >10 nm).

Sample Preparation
Downloaded by [Dalhousie University] at 06:38 20 August 2013

Compared to other techniques, SEM requires little sample preparation. In the


case of nanoparticles, the particles must be collected on a large support medium such
as a filter, grid, or strip of carbon tape. As with TEM, a major concern is that the
collected sample be representative of the original. A recent study by Berube et al.
compared various methods of preparing diesel exhaust particles for SEM analysis.[72]
They compared a three-step method, which involved collection on a filter, sonication
in a solvent in order to prepare a suspension, and then allowing a droplet of
suspension to dry on a standard mount, to directly collecting the particles on a film.
The process of sonicating and wetting the particles was found to cause aggregation
and tighter packing than what was observed for samples collected directly on films.
In order to ensure a representative sample, large numbers of particles should be
collected and analyzed. Furthermore, nonconductive samples may require a coating
of conducting material, such as platinum or carbon, to prevent sample charging.
However, such coatings can affect compositional analyses (since the characteristic
X-rays will also have to travel through the coating) and SE images (most of the
emitted secondary electrons come from the outer few nanometers of material).
The highest resolution SEM images are produced using secondary electrons, and
the ultimate imaging resolution (smallest feature size that may be discerned) in SEM
is limited by the minimum size of the electron beam on the sample (probe size) that
still generates enough SE signal to obtain an image. The best way to increase
ultimate resolution is to use brighter electron sources. An SEM equipped with a
field-emission source can achieve spatial resolutions of 1 nm.
A major advantage of SEM is that it has a large depth-of-field compared to
optical microscopy. This means that for samples with large height variations, a
greater portion of the object will be in focus in a single image. In addition, SEM can
produce images that yield reliable three-dimensional topographic information,
especially if the sample can be imaged at more than one orientation. For example,
Fig. 4 shows two images of complex nanoparticles of lepidocrocite and maghemite,
an iron oxyhydroxide and iron oxide, respectively. While the three-dimensional
morphology of the particles is not obvious in the TEM image, the SEM image
reveals that the particles are shaped like sea urchins. On the other hand, the higher
resolution TEM image demonstrates that the spicules are single crystals with lattice
spacings that conform to that of lepidocrocite (g-FeOOH). Finally, when equipped
with an X-ray detector, SEM can be coupled with EDS, which allows for
simultaneous topographic and elemental mapping.
ORDER REPRINTS

2728 Burleson, Driessen, and Penn


Downloaded by [Dalhousie University] at 06:38 20 August 2013

Figure 4. Electron micrograph images of a mixture of synthetic lepidocrocite and maghemite


nanoparticles on a carbon support film. The left-hand micrograph is an SEM image of the
nanoparticles acquired on a JEOL 6500 FEG SEM using a secondary electron detector. The
right-hand micrograph is a TEM image of a similar group of nanoparticles. The inset in
the upper-right shows a higher magnification TEM image of one of the spine-shaped
structures. The scalebar in the higher magnification TEM image is 10 nm long. Both TEM
images were acquired on a 300 kV FEI CM30 TEM.

A major limitation of SEM is that the electron beam can cause the reduction of
hydrocarbons, which then deposit as a layer of carbonaceous material on the imaged
area that, in turn, reduces image contrast. Thus, it is important that samples are
clean and that preparation materials (e.g., adhesives and epoxies) do not outgas
upon insertion into the microscope. The high vacuum requirements can be detri-
mental to samples that contain volatile components such as water. For inadequately
coated specimens, charging can lead to image distortion, artifacts, or sudden
discharges. Lastly, samples can be damaged or altered upon exposure to the electron
beam.
Hydrated specimens or samples that charge easily but cannot be coated can be
imaged by a newer subset of SEM called environmental SEM (ESEM). In ESEM,
the sample compartment is maintained between 1–10 Torr while the electron source
chamber is maintained at high vacuum by a series of differential pumping stages
along the column. The main disadvantage of ESEM is that its maximum resolution is
only 3–4 nm due to substantial electron scattering by the presence of gas molecules.
Ebert et al. used ESEM to examine the morphological changes and deliquescence of
aerosol particles as a function of relative humidity.[73] They found that ESEM was an
ideal technique for examining the hygroscopic behavior of aerosol particles because,
unlike other available techniques, it allowed them to directly image the aerosol
particles at a range of H2O partial pressures.
Because SEM can accept a wide variety of sample sizes and has a large depth-
of-field, it is an ideal technique for imaging particles ranging in size from 50 nm up
to several mm. While SEM can image nanoparticles smaller than 50 nm, TEM is
ORDER REPRINTS

Characterization of Environmental Nanoparticles 2729

much better suited for imaging such small particles. Recent applications of SEM
(and its complementary technique EDS) in the literature include examining the size,
morphology, and composition of aerosol particles collected from refineries and
urban environments as well as aerosol particles prepared in the laboratory.[74–77]

Particle-Induced X-ray Emission (PIXE)

PIXE is a less commonly used method that is useful for the compositional
analysis of major, minor, and trace constituents of nanoparticles. Trace analysis
(e.g., heavy metals) of atmospheric aerosols can be used to establish the relationship
Downloaded by [Dalhousie University] at 06:38 20 August 2013

between the source (e.g., anthropogenic vs. crustal) and the fate and transport of
such particles.[78] In PIXE, a fixed 2–3 MeV beam of protons or a-particles impinges
on a sample, knocking out core electrons. As in EDS, a characteristic X-ray can be
emitted as the result of an outer shell electron relaxing into the orbital from which
the electron was ejected. PIXE is more bulk sensitive than EDS due to the greater
sample penetration depth of protons and a-particles as compared to electrons.
For a particularly good treatment of the background and application of this
technique the reader is referred to the text of Johansson et al.[79] and two recent
papers that specifically discuss the use of PIXE for the analysis of atmospheric
aerosols.[78,80,81]
While the information obtained from PIXE is very similar to EDS, it has several
advantages over EDS. Most importantly, PIXE spectra do not have the strong
Bremsstrahlung radiation background found in EDS. This makes PIXE an excellent
technique for trace elemental analysis with detection limits on the order of 1015 g
because low intensity X-ray peaks are less obscured by background signal.
The biggest disadvantage of PIXE is that it requires a high-energy particle
accelerator, which means PIXE equipment is large and expensive. Additional
limitations of PIXE are similar to EDS, including the ability to detect only elements
where Z > 4, the low sensitivity to light elements (Z  10), and beam heating of the
sample. Analysis of samples is normally performed under high vacuum to minimize
the scattering of the particle beam by gas molecules. However, for samples that are
large or contain volatile components, it is possible to use a modified PIXE apparatus
to perform external-beam PIXE at atmospheric pressure.[79]
Micro-PIXE is a recent PIXE variation in which a focused particle beam is
rastered across a sample surface to perform elemental mapping.[82–84] Micro-PIXE
can not compete with the spatial resolution of EDS for thin samples. However, for
thicker samples (1 mm), the spatial resolution of the two techniques is comparable
(0.5–1 mm).
PIXE and micro-PIXE are often used for trace elemental analysis of aerosol
particles.[81,85–87] For particles and aggregates that are larger than a micron in dia-
meter, it is possible to perform elemental mapping with micro-PIXE. Unfortunately,
elemental mapping of nanoparticles by micro-PIXE is not possible because its spatial
resolution is larger than the typical particle. This limits the elemental information to
an average of a single isolated particle or group of particles.
ORDER REPRINTS

2730 Burleson, Driessen, and Penn

X-ray Photoelectron Spectroscopy (XPS)

XPS, also known as electron spectroscopy for chemical analysis, is a surface-


senstive technique that is useful for quantitative compositional analysis and the
determination of oxidation state and bonding environment. XPS is extremely surface
sensitive, with the depth probed ranging between two and twenty atomic layers.[31]
This is especially useful when probing nanoparticles that are significantly larger than
a few nanometers in size as the nanoparticle surface can be characterized and the
results compared to bulk analysis results in order to elucidate how the elements are
distributed throughout the particle. It is, however, not suitable for trace analysis.
In XPS, electrons are ejected from the sample as a result of bombardment with
X-rays. The kinetic energies of the ejected electrons and the kinetic energy of the
Downloaded by [Dalhousie University] at 06:38 20 August 2013

incident X-ray photons are used to determine the binding energy of the ejected
elections, which is related to the identity of the elements and their oxidation state and
bonding environment. The ideal sample is conductive, but there are techniques
available for overcoming problems with charging for insulating samples (e.g., low
energy electron flood gun directed at the sample or the use of an internal standard
for which XPS peak positions are known).[88] An example of using XPS to
characterize nanoparticles is recent work by Anschutz and Penn, in which XPS was
used to characterize iron(III) oxyhydroxide nanoparticles, which are common at and
near the Earth’s surface, before and after reaction with a reducing agent. Those
results showed a significant shift in the iron peaks (to weaker binding energies),
which demonstrated that the surface material had been reduced to contain
significant Fe2þ.[89] Lastly, XPS imaging techniques can achieve spatial resolutions
on the order of 1 mm.[90]

CHARACTERIZATION METHODS: BULK SCALE

Methods presented in this section yield ensemble results, reflecting average


information for relatively large (in comparison to the sampling abilities of the above
methods) samples of nanoparticles. The techniques in this section are organized into
chemical composition (atomic spectroscopies), structure (X-ray diffraction and X-
ray absorption and vibrational spectroscopies), and the measurement of magnetic
properties.

Chemical Composition

In addition to the aforementioned techniques (EDS, PIXE, XPS, and EELS),


the average bulk chemical composition of a given sample can be determined using
atomic absorption, atomic emission, and mass spectroscopies. In general, a large
portion of the periodic table can be analyzed using these methods.
AAS (atomic absorption spectrometry), ICP-AES (inductively coupled plasma-
atomic emission spectrometry), and ICP-MS (inductively coupled plasma-mass
spectrometry) share sample preparation and some elements of operation but differ in
terms of the signal detected. In general, a sample solution is atomized at high
ORDER REPRINTS

Characterization of Environmental Nanoparticles 2731

temperature (flame, graphite furnace, or ICP) and either absorption of a mono-


chromatic beam of light, emission at a specific wavelength, or the mass-to-charge
ratio of ions is monitored. With appropriate standards, the intensity of the signal is
used to calculate concentration. Methods of atomizing range from vaporization
followed by atomization of a sample solution to laser ablation of a solid sample.
Sample preparation generally involves the complete or partial dissolution of a
nanoparticulate sample using inorganic acids like nitric and hydrofluoric acids. This
can present some difficulty when working with environmental samples because some
components of the sample may be volatilized during dissolution (e.g., evolution of a
gas upon addition of acid). In addition, some dissolution agents can enrich or deplete
specific elements. For example, hydrofluoric acid should not be used when analyzing
for silicon, as volatile SiF4 is formed during dissolution.[91] On the other hand,
Downloaded by [Dalhousie University] at 06:38 20 August 2013

selective dissolution that results in the concentration of a particular analyte with


respect to the rest of the sample may be advantageous when analyzing for trace
elements.[92–94] Several methods for sample dissolution can be found in the following
references.[91,95–98] For a discussion of laser ablation of solids, the reader is referred to
Ridley[99] and Hill et al.[100] For additional solid sampling techniques see Kurfurst.[101]
In general, AAS and ICP-AES are ideal for the analysis of some main-group
elements and most transition metals. The methods are susceptible to substantial
interferences caused by matrix effects and require that standards closely match the
samples (e.g., solvent, pH, and ionic strength). Typical detection limits for AAS lie in
the range of 10–100 ng/mL, and in general, the sample can be analyzed for only one
element at a time. For a good basic description of AAS see Welz and Sperling,[96]
Van Loon,[95] Dean,[102] or Ebdon et al.[103] In ICP-AES, the instrumental design
usually allows for the simultaneous detection of multiple analytes. Typical detection
limits for ICP-AES lie in the 1–100 ng/mL range. For a detailed discussion of ICP-
AES see Dean[102] or Ebdon et al.[103] In ICP-MS, the mass-to-charge ratio of species
present can be determined with little spectral interference and with superior detection
limits of 1–10 pg/mL. In addition, the concentration of nearly every element of the
periodic table can be quantified using ICP-MS. For a general discussion of the ICP-
MS technique see Dean,[102] Ebdon et al.,[103] or Jarvis et al.[104] For a few specific
examples of environmental analysis using ICP-MS see Hill et al.,[100] Balaram,[97] or
Gomez et al.[105] For a good comparison of AAS, ICP-AES and ICP-MS techniques,
see Dean,[102] Balaram,[97] or Skoog et al.[106]
Some advantages of using AAS, ICP-AES, and ICP-MS for nanoparticle
characterization are: (1) the methods are relatively quick and inexpensive; (2) the
methods have good detection limits; and (3) average composition of a sample of
nanoparticles is obtained. Some disadvantages are: (1) the techniques are destructive;
(2) sample preparation and matrix effects can cause major artifacts; and (3) the
average composition of the sample is obtained. ICP-AES and ICP-MS instruments
are less common and much more expensive than AAS. A major advantage of ICP-
AES and ICP-MS over AAS is that multiple elements can be analyzed
simultaneously whereas in AAS, elements are typically analyzed sequentially or, at
best, two to three elements simultaneously. Lastly, AAS and AES are good for
transition metals and some main-group elements while ICP-MS has the added
advantage of being able to quantify nearly every element of the periodic table and
can even quantify the concentration of isotopes.
ORDER REPRINTS

2732 Burleson, Driessen, and Penn

Structure

X-ray Diffraction (XRD)

X-ray diffraction is used most often to identify crystalline phases by comparison


to a known database, measure atomic spacings (with the use of appropriate
standards), measure structural properties of the crystalline phases present (e.g.,
preferred orientation and particle size), and to probe atomic arrangements in
amorphous materials. The identification of crystalline materials is often aided by
chemical analysis in order to identify the elements present. The mass fractions of the
phases present in a sample can be determined by comparison to known standards.
Since XRD is relatively quick and is usually a nondestructive technique, it is often
Downloaded by [Dalhousie University] at 06:38 20 August 2013

one of the first methods employed in the characterization of nanoparticles.


When the X-ray beam interacts with an ordered array of atoms, a diffracted
beam is observed when the wavelength and angle of incidence with respect to the
crystal result in constructive interference of the diffracted X-rays, as described by
Bragg’s Law (see Electron Diffraction section). The main factors that determine
peak position and intensity in XRD patterns of crystalline materials is the size and
shape of the unit cell and the atomic weight and positions of the atoms in the unit
cell. Thus, XRD serves as a probe of crystal structure. Unfortunately, XRD cannot
be used to directly identify the elements present. For a detailed treatment of the
instrumentation and theory of XRD, the reader is referred to Cullity and Stock[107]
or Toney.[108] A typical XRD instrument consists of an X-ray source, optics, sample,
and a detector. The detector is often scanned through 2y while the sample is scanned
through y, which is why the intensity data from XRD is typically plotted as a
function of degrees 2y.
The Powder Diffraction File, which is maintained by the International Center
for Diffraction Data, is a database of tens of thousands of XRD powder patterns. By
comparing an experimental pattern to this database, a qualitative identification of
the crystalline phases present can be performed. If the user has some additional
information about the sample (e.g., collection site, synthesis conditions, chemical
composition), identification can be quick and straightforward. However, it is critical
that the user evaluate the database matches in the context of the sample analyzed
(e.g., compositional analysis results, synthetic conditions, or field site from which the
particles were collected). Database matches do not necessarily mean that material is
present in the sample analyzed.
In addition to identifying a crystalline material, XRD can be used to determine
the average size of nanocrystals and is practical for average particle sizes
below 50 nm. The width of a diffraction peak is inversely related to the size of
the diffracting crystallite. The full width at half maximum (FWHM) value for the
diffraction peaks can be used to calculate the average crystallite size along a specific
crystallographic direction.[109,110] More advanced treatments of peak shape and
width can yield some information about particle size distribution.[111] Specific
considerations for dealing with nanoparticles in XRD are discussed in Huber[109] and
Zanchet et al.[112]
Nanoparticle samples are typically mounted on a sample plate in powder form.
XRD provides average information for a sample, so it is critical that the sample be
ORDER REPRINTS

Characterization of Environmental Nanoparticles 2733

representative of the material under investigation. In general, X-ray diffraction


experiments require sample masses on the order of tens of milligrams. Ideally, the
sample preparation method (e.g., slurry, cavity, or dusting mounts) should result in
random orientation or reproducible preferred orientation of sample particles.[113]
Preferred orientation results in intensity ratios that differ from those observed for
randomly oriented samples, and intensity ratios are often important to the
characterization of nanoparticles. Practically speaking, a thin and smooth layer of
particles on a planar mount (often a glass slide) is often sufficient to give good
results. Finally, for crystallite size analysis or the characterization of amorphous or
semicrystalline materials, a single crystal plate is often employed in order to reduce
the background signal due to diffraction from the mount. Some special considera-
tions in sampling and preparing minerals are discussed in Blanton et al.[114]
Downloaded by [Dalhousie University] at 06:38 20 August 2013

The major advantages of XRD are: (1) crystalline materials can be identified
quickly; (2) amorphous and crystalline materials can be distinguished; (3) XRD
patterns are additive; (4) the method is quick and relatively inexpensive; (5) particle
size and size distribution information can be obtained for crystalline particles; and
(6) it’s possible to perform the analysis using controlled atmosphere chambers (e.g.,
anaerobic conditions or controlled humidity). Major weaknesses include: (1) trace
components (< 1 mass %) can be difficult to detect; (2) elements cannot be directly
identified; (3) the pattern is an average over all particles in the scattering volume; and
(4) peak shapes can be difficult to interpret.

X-ray Absorption Spectroscopy (XAS)

Generally speaking, X-ray absorption spectroscopy (XAS) methods involve


irradiating a sample with a tunable high intensity beam of X-rays (i.e., from a
synchrotron source) and monitoring the intensity of the beam after interacting with
the sample. When the energy of the X-rays is sufficient to eject a core electron, an
absorption edge is observed in the spectrum. Peak shapes and positions and
oscillations in post-edge intensity are sensitive to local order and the identity,
arrangement, and bonding of atoms around the absorbing element. Three types of
XAS will be discussed here: EXAFS (extended X-ray absorption fine structure),
XANES (X-ray absorption near edge structure; also known as NEXAFS), and
SEXAFS (surface extended X-ray absorption fine structure).
EXAFS and XANES probe structure, interatomic distance, coordination
number, and degree of disorder with respect to one element, out to a distance of
6 Å from the absorbing atom. The absorption coefficient of the material is
calculated and plotted as a function of incident X-ray energies.[112,115,116] EXAFS
refers to the portion of the spectrum above 40 eV beyond the absorption edge. When
the energy of the incident X-rays exceeds the binding energy of core electrons in the
absorbing element, electrons are ejected and subsequently backscattered from
neighboring atoms. Constructive or destructive interference between the X-rays and
the backscattered electrons gives rise to oscillations in the spectra that are related to
the average identity, number, and arrangement of the absorbing element’s nearest
neighbor atoms.[112,116,117] In addition, EXAFS spectra can be used as fingerprints
for the identification of unknown materials, and the application of a least-squares
ORDER REPRINTS

2734 Burleson, Driessen, and Penn

fitting protocol can provide quantitative measurements of both the type and
percentage of phases present. Recent work has used EXAFS to characterize hydrous
iron oxide nanoparticles, which are important materials in the biogeochemical
cycling of chemical species in the environment, formed in the presence of zinc[118] and
arsenate.[119] Figure 5 shows Fe K-edge EXAFS spectra, and the corresponding
Fourier transforms, of iron oxyhydroxide nanoparticles aged at 90 C from 0–25
days. Comparison of the experimental spectra to reference spectra and the observed
Downloaded by [Dalhousie University] at 06:38 20 August 2013

Figure 5. Fe K-edge EXAFS spectra (a) and corresponding Fourier transforms (b) of iron
oxyhydroxide nanoparticles aged at 90 C for 0–25 days. On the left-hand figure, the x-axis (k)
represents energy converted to momentum space following the normalization of the EXAFS
data to a fixed point in energy space, and the y-axis (C(k)k3) is a k-cubed weighted expression
of the EXAFS function. The Fourier transforms are similar to radial distribution functions
and have as their x-axis the radial distance (þ phase shift) in Angstroms and as the y-axis the
Fourier transform peak magnitude, which is a function of the coordination number, identity
of scattering neighbors, and degree of structural disorder as identified by the Debye-Waller
factor. Model compound spectra are also presented (labeled ferrihydrite and goethite) to show
the structural change from ferrihydrite to goethite as a function of aging time. This is
demonstrated by changes in the position or intensity of various EXAFS features (two sets of
peaks are highlighted by grey vertical guidelines). An increase in the amplitude of Fourier
transform features between 2.2–3.2 Å (Fe–Fe interactions) also accompanies the transition
from ferrihydrite to goethite. Figure courtesy of C. Kim and G. Waychunas at Berkeley
National Laboratories.
ORDER REPRINTS

Characterization of Environmental Nanoparticles 2735

shifts in peak intensities, position, and shape as a function of aging time demonstrate
that ferrihydrite transforms to goethite over time.
Increasingly bright synchrotron sources can now conduct spatially resolved
EXAFS analysis with spot sizes as small as 1  1 mm. While this is not yet small
enough to effectively map nanoparticles, it does minimize the total amount of sample
needed to collect a usable EXAFS spectrum. For example, for pure phases (e.g., the
iron oxyhydroxide samples highlighted in Fig. 5) 5–10 mg of sample is sufficient.
XANES refers to the portion of the spectrum with X-ray energies ranging from
just below (pre-edge) to just above (post-edge) the absorption edge and is a sensitive
probe of valence and site symmetry for many elements. It is more sensitive but less
quantitative than EXAFS. Pre-edge features are the result of electronic transitions
and provide information about valence and bonding environment of the absorbing
Downloaded by [Dalhousie University] at 06:38 20 August 2013

element. Post-edge features near the absorption edge provide information about the
symmetry around the absorbing element.[112,116,117] Figure 6 shows Cr K-edge
XANES spectra for Cr(III) and Cr(VI) species,[116] illustrating how XANES spectra
can be used to determine the oxidation state of the absorbing element. SEXAFS is a
surface sensitive technique that provides information about nanoparticle surfaces
and the way adsorbed species are bonded to surfaces. In SEXAFS, the X-ray
absorption coefficient is usually determined indirectly by monitoring fluorescence,
Auger electron emission, or total or partial electron yield. For a detailed treatment of
the theory and application of SEXAFS, the reader is referred to Cheng et al.,[120]
Lagarde,[121] Norman,[122] and Stor.[123]

Figure 6. Cr K-edge XANES specta for Cr(III) and Cr(VI) species, illustrating how
NEXAFS spectra can be used to determine the oxidation state of the absorbing element.
Reprinted with permission from Fendorf et al.[116]
ORDER REPRINTS

2736 Burleson, Driessen, and Penn

XAS methods can provide detailed structural information about environmental


nanoparticles. The methods are particularly useful for semi-crystalline and amo-
rphous nanoparticles (such as ferrihydrite, a hydrous iron oxyhydroxide nanoma-
terial that is common on and near the Earth’s surface) and are often used to
characterize surficial materials that contain substantial amounts of nanoparticulate
components (e.g., soils[116]). In this regard, XAS methods fill an important niche,
especially when compared to XRD, which is most suitable for the characterization of
crystalline nanoparticles. Major advantages include: (1) in situ experiments are
possible since high vacuum conditions are not required; (2) it is theoretically possible
to distinguish between surface-bound and bulk chemical species; (3) elements and
oxidation states can be directly identified; (4) order and disorder can be char-
acterized; (5) the detailed arrangement of atoms around the absorbing atom can be
Downloaded by [Dalhousie University] at 06:38 20 August 2013

determined; (6) surface sensitive measurements can be obtained; and (7) hetero-
geneous samples can be characterized. Major disadvantages include: (1) synchrotron
radiation is required; (2) data interpretation is complicated; and (3) results are an
average over all particles present.

Vibrational Spectroscopy

Raman and infrared (IR) spectroscopies are vibrational spectroscopies


commonly used to identify materials by comparison to knowns, obtain information
about how chemical species are bound to one another in a sample, and quantify the
concentration of specific chemical species in a sample. These vibrational spectro-
scopies are usually nondestructive and can sometimes be performed in situ. While no
vibrational spectroscopy is inherently surface sensitive, the small size of nanopar-
ticles means that vibrational spectroscopy can be used to characterize surface-bound
species.
In IR spectroscopy, the transmission of infrared radiation through a sample is
monitored. Absorption occurs when the frequency of the infrared radiation matches
the energy required to excite a vibrational mode in the sample and a change in dipole
moment accompanies the vibration. The frequency of a peak, or band, in the
infrared spectrum is used to identify bonds and determine bond strengths between
specific groups of atoms. IR spectra can be used as fingerprints for the identification
of materials by comparison to knowns, and IR spectroscopy can be used to quantify
chemical species. Fourier transform IR (FTIR) instruments, which convert a
spectrum collected in the time domain to the frequency domain, allow the entire
spectrum to be collected at one time, making it possible to collect a large number of
spectra in a short time, which results in a substantial improvement in signal-to-noise
ratios. FTIR spectrometers are common and are generally much faster and more
sensitive than dispersive instruments. Samples must be infrared transparent and
allow enough of the beam to pass through to take a measurement. For a detailed
discussion of infrared spectroscopy see Ingle et al.[124] or Skoog et al.[106] Specific
applications of infrared spectroscopy to nanoparticles can be found in Schwertmann
et al.[125] or Woolridge et al.[126]
Raman spectroscopy typically uses an intense monochromatic beam of light
(usually from a laser) to excite the sample, and upon relaxation, light is reradiated in
ORDER REPRINTS

Characterization of Environmental Nanoparticles 2737

all directions. Raman spectrometers usually monitor light emitted in a direction


perpendicular to the laser beam. The vast majority of the reradiated light has the
same wavelength as the incoming radiation (Rayleigh scattering) and is the result of
the sample relaxing back to the same vibrational energy level from which it was
excited. When the sample relaxes down to a different vibrational energy level
(Raman scattering), the emitted photon has an energy that is equal to the incident
energy plus the energy change associated with the change in vibrational state (can be
positive or negative). Raman active vibrational modes are those that are
accompanied by a change in polarizability of the molecule.
Raman spectroscopy is inherently an insensitive technique, with less than one
out of 105 photons actually undergoing the Raman process. Fourier transform
instruments obtain results with substantially improved signal-to-noise ratios. Fur-
Downloaded by [Dalhousie University] at 06:38 20 August 2013

thermore, fluorescence from the sample, which is the result of electronic transitions
caused by the absorption of incident photons, can complicate data collection. In
general, the use of a lower frequency source can eliminate fluorescence. For general
discussions of Raman spectroscopy see Ingle et al.[124] or Skoog et al.[106] For specific
applications of Raman spectroscopy to solid materials see Lyon et al.,[127]
Pelletier,[128] Weber and Merlin,[129] or Mulvaney and Keating.[130] Specific
applications of Raman spectroscopy to environmental particles include works by
Tarassov et al.[110] and de Faria et al.[131] Tarrasov examined the incorporation of
tungsten in environmental goethite and hematite nanoparticles from Bulgaria in
order to elucidate the mechanism by which tungsten is incorporated into these
mineral particles.[110] De Faria used Raman spectroscopy in order to characteriza-
tion iron oxide and oxyhydroxide phases.[131]
The two major differences between IR and Raman spectroscopy lie in the exci-
tation source, the signal monitored, and the selection rules for observed vibrational
modes as mentioned above. Some IR-active vibrational modes are not Raman-active,
and not all vibrational modes are active for even one of the two techniques. For
instance, the strong absorption of infrared radiation by water makes IR difficult to use
for the characterization of wet samples. Raman spectroscopy, on the other hand, is
not sensitive to the presence of water and therefore well-suited to monitor reactions
that occur in aqueous environments. Lastly, Raman is better suited for characterizing
lower energy vibrational modes such as metal-oxygen bonds, whereas IR is better
suited for higher energy vibrational modes such as carbon-hydrogen bonds.
Lastly, while no vibrational spectroscopy is inherently surface sensitive, the large
surface area-to-volume ratios of nanoparticles means that surface-bound species can
often be characterized, especially if the species of interest occurs only at the surface.
These can either be the species that terminate the bulk structure, for example,
hydroxyl groups at the surface of many metal oxides, or adsorbed species. For
examples of infrared spectroscopy used to characterize surface bound species, see
work by Driessen et al.[132–134] In addition, Khaleel and Dellinger used FTIR to
characterize nanophase alumina catalyst that was used in the chemical transforma-
tion of carbon tetrachloride. Their results showed that there was a significant
concentration of surface hydroxyls before and after heat treatment. In addition, they
used the method to monitor the concentrations of the gaseous products of the
carbon tetrachloride transformations.[135]
ORDER REPRINTS

2738 Burleson, Driessen, and Penn

Magnetic Measurements

Magnetic measurements can be used to quickly and efficiently characterize iron-


rich nanoparticles such as the iron oxides and oxyhydroxides, which are ubiquitous
on and near the Earth’s surface.[136,137] Rancourt does an excellent job of discussing
the theory and application of these methods.[136] Magnetometry is a collection of
techniques that probe the magnetism of materials using one of many possible
instrument designs.[138] For example, the temperature at which magnetic transitions
occur is sensitive to oxidation state, crystal structure, and particle size. Thus, moni-
toring magnetization as a function of temperature can serve to identify materials and
determine particle size. These methods can be used to identify magnetic nano-
particles (e.g., distinguish between goethite, hematite, maghemite, and magnetite) in
Downloaded by [Dalhousie University] at 06:38 20 August 2013

trace amounts and even when extremely heterogeneous mixtures of nanoparticles are
analyzed.
Mössbauer, or g ray spectroscopy, is a nondestructive technique that probes
transitions in nuclei in order to characterize magnetic properties and obtain
information on the oxidation state, coordination number, and bonding environment
around the absorbing atom. It is also used to identify magnetic components in
heterogeneous samples. Monochromatic g rays are used to irradiate the sample, and
the incident g ray energy is varied by translating the source (i.e., the Doppler shift).
The transmitted radiation is monitored, and absorption is plotted vs. the velocity of
the source translation (positive for towards the source and negative for away from
the source).[136,139,140] One limitation of Mössbauer spectroscopy is that the analysis
is most commonly limited to Fe and Sn due to availability of appropriate g ray
sources. Several examples of using Mössbauer spectroscopy include the character-
ization of iron containing aerosols[139] and sediments.[125]

CHARACTERIZATION METHODS: PARTICLE SIZING

The determination of particle size and size distribution of nanoparticles is often


a critical characterization because many properties of particles are size dependent. A
wide range of methods is available, ranging from the size determination of one
particle at a time (e.g., via electron microscopy) to the determination of the average
size of many particles at a time (e.g., XRD peak broadening analysis). Four
additional methods not treated heretofore that yield particle size information are
Brunauer Emmett Teller (BET) surface area, photon correlation spectroscopy,
aerosol time-of-flight mass spectrometry, and differential mobility analysis.

Average Particle Size Determination

Brunauer Emmett Teller (BET) Surface Area

The BET surface area method measures the volume of gas required to adsorb a
single monolayer of gas molecules (or atoms) to the accessible surface area of a given
mass of sample.[71,141] Usually, nitrogen is the gas of choice. The total accessible
surface is calculated using the BET equation, which was named for and developed by
ORDER REPRINTS

Characterization of Environmental Nanoparticles 2739

Brunauer, Emmett, and Teller and which relates the monolayer volume of gas
adsorbed, the size of the gas molecules, and the mass of the sample to the total
surface area per mass.[142]
The BET surface area method is inexpensive and relatively easy to perform, and
the method directly and quickly measures the accessible surface area per mass of
sample. Furthermore, it requires little sample preparation other than to ensure that
the sample particles are dry, degassed, and well-separated before performing
measurements. If particles are not adequately degassed and well-separated, surface
area measurements will be artificially low. Unfortunately, no information about the
size and shape distributions is obtained. In addition, more complicated data analysis
enables the characterization of porous materials. When used in combination with
another method, such as electron microscopy, a description of particle size, size
Downloaded by [Dalhousie University] at 06:38 20 August 2013

distribution, and accessible surface area can be produced. Recent studies have used
BET surface area measurements to examine the surface area change of iron oxide
nanoparticles during phase changes[143] and the calcination of tin oxide and titanium
dioxide nanoparticles.[144,145]

Photon Correlation Spectrometry (PCS)

Photon correlation spectrometry is a dynamic light scattering (DLS) method


often used to measure particle size and size distribution for samples suspended in a
liquid. A detailed discussion of the technique can be found in several sources.[146–151]
In a common PCS set-up, a suspension of particles is placed in the path of a
monochromatic, highly-coherent light source provided by a laser, and the intensity
of scattered light is measured by a photo-electric detector, such as a photomultiplier
tube or a photodiode array. Since the suspended particles are in constant Brownian
or thermal motion, the intensity of the scattered light fluctuates with time. An
analysis of this time-dependent fluctuation using an autocorrelator gives an
autocorrelation function whose delay time is directly related to the diffusion
coefficients of the particles. The hydrodynamic radius of the particles can be
estimated from the diffusion coefficients using the Stokes-Einstein equation. This
technique works best on monodisperse suspensions of 5–3000 nm particles. For a
polydisperse sample of particles, more than one delay time will be observed, which
means that particle size distribution information can be obtained. However, the data
analysis is challenging, and such results must be confirmed using another method
(e.g., electron microscopy or XRD).
Advantages of PCS include that it can determine particle size of monodisperse
and polydisperse samples in situ and in real-time with generally little to no sample
preparation. The particles must be suspended in a liquid, which can be a major
advantage if studying particle size or aggregation in suspensions. However, several
factors can result in inaccurate particle size values. If the particle concentration is too
high, photons are scattered multiple times before they exit the suspension and are
detected, which generally leads to an underestimation of the particle size. On the
other hand, if the particle concentration is too low, fluctuations in the concentration
of particles in the scattering volume with time significantly contributes to the inten-
sity fluctuation, which leads to inaccurate particle size determinations. For large
ORDER REPRINTS

2740 Burleson, Driessen, and Penn

particles (>500 nm) it can be difficult, if not impossible, to find a concentration


where multiple scattering or intensity fluctuations caused by a small number of par-
ticles in the sampling volume do not lead to poor estimates of particle size. Finally, if
the suspension has strong interparticle interactions, the diffusion coefficient can vary
with concentration, which also leads to an inaccurate determination of particle size.
Several recent studies have used PCS to examine environmentally-relevant
particles, including the size distribution of iron oxide nanoparticles, magnesium
oxide nanoparticles, and soot particles.[152–154] PCS is most effective for nanopar-
ticles that are easily suspended, roughly spherical, and do not interact strongly with
each other. PCS is also a good choice when specifically examining particle size and
aggregation in suspension. The technique is particularly useful when combined with
a technique like electron microscopy. For example, if a sample of particles is poly-
Downloaded by [Dalhousie University] at 06:38 20 August 2013

disperse, electron microscopy results can produce a small particle size distribution
dataset that can be used as a starting model for the distribution determined from the
autocorrelation function collected using DLS. In addition, if the particles appear
aggregated in the electron microscope, DLS could be used to determine whether the
particles might also have been aggregated in solution.

Single Particle Size Determination

Aerosol Time-of-flight Mass Spectrometry (ATOFMS)

Aerosol time-of-flight mass spectrometry is a relatively new technique capable of


measuring the size and composition of individual airborne nanoparticles. The
instrument samples airborne particles by keeping the interior of the instrument at a
lower pressure than ambient, which drives the flow of ambient air through a nozzle
and into the instrument and results in the acceleration of the particles to their
terminal velocity. That velocity can be converted to aerodynamic diameter because
terminal velocity is particle size dependent. A desorption/ionization laser is timed to
fire when the particle reaches the appropriate location and results in the production
of both positive and negative ions, which are then sent to a dual polarity mass
spectrometer that classifies the ions according to mass-to-charge ratio. The
instrument is capable of sampling several particles per second, which means that
the data produced includes the size and chemical composition of each nanoparticle
analyzed. For details on the construction and operation of an ATOFMS, the reader
is referred to Fergenson et al.,[155] Gard et al.,[156] or Hughes et al.[157] Particle size
distributions can be skewed because particle detection efficiency is size dependent,
with the largest fraction being sampled most efficiently.[25,158] Furthermore,
nanoparticle characteristics can be monitored as a function of time (e.g., monitoring
the size distribution and chemical composition of airborne nanoparticles in urban
areas as a function of time[156,157,159–163]).

Differential Mobility Analysis (DMA)

A differential mobility analyzer (DMA) can be used to measure the particle size
distribution of an aerosol sample and to size-select aerosol particles for subsequent
ORDER REPRINTS

Characterization of Environmental Nanoparticles 2741

analysis.[164–166] It operates by first passing a stream of particles suspended in a gas


through a neutralizer, a device that imparts to the particles a Boltzmann distribution
of charges. The particles, along with a laminar flow of extra gas, then travel through
an annulus space created by an outer tube and central rod. The outer tube is
grounded while the voltage on the central rod can be varied across a wide range of
potentials. Only particles with a narrow range of differential mobilities will exit
through the exit gap for a given rod potential, while the other particles will deposit
on the walls or exit through an excess air flow port. Particles that pass through the
exit gap are normally detected using a condensation particle counter (CPC), which is
also often referred to as a condensation nuclei counter. Particle size is then calculated
from the differential mobilities.
The major advantages of the DMA are that it is simple and cheap, and particles
Downloaded by [Dalhousie University] at 06:38 20 August 2013

with a wide range of diameters (2–1000 nm) can be analyzed in real time with little to
no sample preparation. In addition, samples that are already aerosols require little to
no sample preparation except the addition of a carrier gas to dilute the particles and
carry them through the DMA. Particle suspensions and dry powders can also be
analyzed. Particles suspended in volatile liquids can be converted into solid aerosol
particles by nebulizing the liquid suspension and then drying the resulting droplets
in a carrier gas. A wide range of powdered materials can be converted into aerosols
by various dry dispersion methods.[165] Lastly, a DMA can size-select particles
for analysis using other instrumentation by directing the exit flow to a second
instrument (e.g., mass spectrometer) or by collecting the particles on a support
medium.
The primary limitations of a DMA are that the samples must be aerosols with
concentrations between 1 and 108 particles cm3 and that DMA does not directly
provide chemical or structural information. In addition, as particle size decreases, a
smaller percentage of particles are charged by the neutralizer. Therefore, the DMA
has a lower sensitivity to smaller particles because only charged particles with the
correct mobility leave the device through the exit slit for a particular potential
setting. Finally, the particle diameter measured by the DMA is a mobility diameter,
which is equivalent to the diameter of a spherical particle with the same surface area
as the particle analyzed. Since particles are often irregularly shaped, the actual
particle size and mobility diameter are frequently quite different. Thus, the DMA by
itself gives no direct information about particle shape or aggregation.
When a DMA is used to size-select particles, exiting particles can either be
collected for further analysis (e.g., electron microscopy) or passed directly into
another analytical instrument. In work by Weber et al., silver particle agglomeration
was studied by DMA, ICP-AES, SEM, and TEM.[167] In their work, DMA size-
selected particles were sent directly into an ICP-AES to determine the amount of
silver (in the form of particle agglomerates) exiting the DMA. A portion of the size-
selected particles were also collected on a filter and then analyzed by SEM and TEM
to determine particle morphology. The DMA and the various techniques that can be
coupled to it have many direct applications to nanoparticles, especially atmospheric
nanoparticles. DMAs have been coupled to mass spectrometers in order to obtain
simultaneous size and chemical information of soot nanoparticles.[168] Tandem
DMA (TDMA) systems send a stream of particles through a DMA to size-select
the particles and then direct those size-selected particles through a reaction zone
ORDER REPRINTS

2742 Burleson, Driessen, and Penn

(with heating or oxidizing conditions) and into a second DMA in order to determine
the final particle size distribution. TDMA systems have been used to study the
hygroscopicity and oxidation of soot and other atmospheric aerosol particles by
measuring particle size changes as a result of heat treatment in the presence of a
variety of gasses (e.g., inert gas vs. oxygen).[169–172]

CONCLUSION

In general, nanoparticle characterization employs multiple techniques. For


example, Higgins et al. used a TDMA system to examine soot nanoparticles and how
Downloaded by [Dalhousie University] at 06:38 20 August 2013

particle mobility diameter changes as a function of heating.[173] TEM revealed that


these particles have an open architecture and are aggregates of smaller primary
particles. The DMA results of particles before and after heating show that in the
absence of an oxidizing gas (like O2) the particles became only slightly smaller
through either a structural rearrangement of the particles or the evaporation of
semivolatile components. When the particles were heated in the presence of O2, they
became substantially smaller due to particle oxidation. In another example,
Weinbruch et al. characterized aerosol particles emitted from a nickel refinery
using SEM, EDS, and TEM in order to determine the size, shape, composition, and
phase of individual particles.[75] SEM results showed a range of nanoparticle sizes
and shapes, and EDS was used to identify those nanoparticles as metal oxides, metal
sulfides, and soot. Coupling SEM and EDS results for each nanoparticle analyzed
showed that the soot particles typically had an agglomerate chain-like structure
while many of the metal oxide particles were larger and spherically shaped.
Furthermore, TEM showed most particles were a mixture of several phases. Such
data is indispensable when it comes to predicting and understanding the fate,
transport, and environmental and health impacts of such nanoparticles.

ACKNOWLEDGMENTS

Parts of this work were supported by the University of Minnesota, the US


Department of Energy Environmental Management Sciences Program, and the
DOE Office of Basic Energy Science, Chemistry Division. Some of the data
presented were collected at the Environmental Molecular Sciences Laboratory, a
DOE User Facility operated by Battelle for the DOE Office of Biological and
Environmental Research at Pacific Northwest National Laboratory.

REFERENCES

1. Anastasio, C.; Martin, S.T. Atmospheric nanoparticles. Rev. Miner. Geochem.


2001, 44, 293–349.
ORDER REPRINTS

Characterization of Environmental Nanoparticles 2743

2. Cass, G.R.; Hughes, L.A.; Bhave, P.; Kleeman, M.J.; Allen, J.O.; Salmon, L.G.
The chemical composition of atmospheric ultrafine particles. Philos. Trans. R.
Soc. London, Ser. A 2000, 358 (1775), 2581–2592.
3. Banfield, J.F.; Zhang, H. Nanoparticles in the environment. Rev. Miner.
Geochem. 2001, 44, 1–58.
4. Cassee, F.R.; Muijser, H.; Duistermaat, E.; Freijer, J.J.; Geerse, K.B.;
Marijnissen, J.C.M.; Arts, J.H.E. Particle size-dependent total mass deposition
in lungs determines inhalation toxicity of cadmium chloride aerosols in rats.
Application of a multiple path dosimetry model. Arch. Toxicol. 2002, 76 (5–6),
277–286.
5. Brown, D.M.; Wilson, M.R.; MacNee, W.; Stone, V.; Donaldson, K. Size-
dependent proinflammatory effects of ultrafine polystyrene particles: a role for
Downloaded by [Dalhousie University] at 06:38 20 August 2013

surface area and oxidative stress in the enhanced activity of ultrafines. Toxicol.
Appl. Pharmacol. 2001, 175 (3), 191–199.
6. Anderson, H.R. Differential epidemiology of ambient aerosols. Philos. Trans.
R. Soc. London, Ser. A 2000, 358 (1775), 2771–2785.
7. Donaldson, K.; Stone, V.; Clouter, A.; Renwick, L.; MacNee, W. Ultrafine
particles. Occup. Environ. Med. 2001, 58 (3), 211–216.
8. Lien, H.L.; Zhang, W.X. Nanoscale iron particles for complete reduction of
chlorinated ethenes. Colloid. Surface. A 2001, 191 (1–2), 97–105.
9. Zhang, W.X.; Wang, C.B.; Lien, H.L. Treatment of chlorinated organic
contaminants with nanoscale bimetallic particles. Catal. Today 1998, 40 (4),
387–395.
10. Zhang, W.X. Nanoscale iron particles for environmental remediation: an
overview. J. Nanopart. Res. 2003, 5 (3–4), 323–332.
11. ATSDR. Toxicological Profile for Chromium; US Department of Health and
Human Services, Public Health Service: Atlanta, GA, 2000.
12. Rowbotham, A.L.; Levy, L.S.; Shuker, L.K. Chromium in the environment: an
evaluation of exposure of the UK general population and possible adverse
health effects. J. Toxicol. Env. Heal. B 2000, 3 (3), 145–178.
13. Murray, C.B.; Kagan, C.R.; Bawendi, M.G. Self-organization of CdSe
nanocrystallites into 3-dimensional quantum-dot superlattices. Science 1995,
270 (5240), 1335–1338.
14. Alivisatos, A.P. Semiconductor clusters, nanocrystals, and quantum dots.
Science 1996, 271 (5251), 933–937.
15. Alivisatos, A.P. Perspectives on the physical chemistry of semiconductor
nanocrystals. J. Phys. Chem. 1996, 100 (31), 13226–13239.
16. Efros, A.L.; Rosen, M. The electronic structure of semiconductor nanocrystals.
Annu. Rev. Mater. Sci. 2000, 30, 475–521.
17. Navrotsky, A. Thermochemistry of nanomaterials. Rev. Miner. Geochem.
2001, 44, 73–103.
18. Buffat, P.; Borel, J.P. Size effect on the melting temperature of gold particles.
Phys. Rev. A: At. Mol. Opt. Phys. 1976, 13, 2287–2298.
19. Penn, R.L.; Anschutz, A.J.; Burleson, D.; Mostrom, A.P.; Guyodo, Y.;
Banerjee, S. Abstracts of Papers, 225th National Meeting of the American
Chemical Society, New Orleans, LA, March 23–27, 2003; American Chemical
Society: Washington, DC, 2003; GEOC-178.
ORDER REPRINTS

2744 Burleson, Driessen, and Penn

20. Rao, C.N.R.; Kulkarni, G.U.; Thomas, P.J.; Edwards, P.P. Size-dependent
chemistry: properties of nanocrystals. Chem. Eur. J. 2002, 8 (1), 28–35.
21. Guczi, L.; Petoe, G.; Beck, A.; Frey, K.; Geszti, O.; Molnar, G.; Daroczi, C.
Gold nanoparticles deposited on SiO2/Si(100): correlation between size,
electron structure, and activity in co oxidation. J. Am. Chem. Soc. 2003,
125 (14), 4332–4337.
22. Hochella, M.F.; Banfield, J.F. Chemical Weathering Rates of Silicate Minerals.
Reviews in Mineralogy; Mineralogical Society of America: Washington, DC,
1995; Vol. 31, 353–406.
23. Zhang, H.Z.; Gilbert, B.; Huang, F.; Banfield, J.F. Water-driven structure
transformation in nanoparticles at room temperature. Nature 2003, 424 (6952),
1025–1029.
Downloaded by [Dalhousie University] at 06:38 20 August 2013

24. Penn, R.L.; Burleson, D.J. Two-step growth of goethite nanoparticles from
ferrihydrite, in prep.
25. Allen, J.O.; Fergenson, D.P.; Gard, E.E.; Hughes, L.S.; Morrical, B.D.;
Kleeman, M.J.; Gross, D.S.; Gaelli, M.E.; Prather, K.A.; Cass, G.R. Particle
detection efficiencies of aerosol time of flight mass spectrometers under ambient
sampling conditions. Environ. Sci. Technol. 2000, 34 (1), 211–217.
26. Penn, R.L.; Zhu, C.; Xu, H.; Veblen, D.R. Iron oxide coatings on sand grains
from the atlantic coastal plain: high-resolution transmission electron micro-
scopy characterization. Geology 2001, 29 (9), 843–846.
27. Williams, D.B.; Carter, C.B. Transmission Electron Microscopy: A Textbook for
Materials Science; Plenum Press: New York, 1996.
28. Buseck, P.R.; Cowley, J.M.; Eyring, L. High-Resolution Transmission Electron
Microscopy and Associated Techniques; Oxford University Press: New York,
1988.
29. Fultz, B.; Howe, J.M. Transmission Electron Microscopy and Diffractometry of
Materials; Spring-Verlag: Berlin, New York, 2001.
30. Wang, Z.L. Transmission electron microscopy of shape-controlled nanocrystals
and their assemblies. J. Phys. Chem. B 2000, 104 (6), 1153–1175.
31. Geiss, R.H EDS: Energy-Dispersive X-ray Spectroscopy. In Encyclopedia of
Materials Characterization; Brundle, R.C., Evans, C.A., Wilson, S., Eds.;
Butterworth-Heinemann: Boston, 1992; 120–134.
32. Joy, D.C.; Alton, D.; Romig, J.; Goldstein, J.I. Principles of Analytical Electron
Microscopy; Plenum Press: New York, 1986.
33. Garratt-Reed, A.J.; Bell, D.C. Energy Dispersive X-Ray Analysis in the Electron
Microscope; Bios Scientific Publishers, Ltd.: Oxford, U.K., 2003.
34. Egerton, R.F. Electron Energy-Loss Spectroscopy in the Electron Microscope;
Plenum Press: New York, 1996.
35. Ahn, C.; Disko, M.; Fultz, B. Transmission Electron Energy Loss Spectrometry
in Materials Science; Times of Acadiana Pr Inc., 1992.
36. Lichte, H. Electron image plane off-axis holography of atomic structures. In
Advances in Optical and Electron Microscopy; Mulvey, T., Sheppard, C.J.R.,
Eds.; Academic Press: New York, 1991; Vol. 12, 25.
37. Orchowski, A.; Rau, W.D.; Lichte, H. Electron holography surmounts
resolution limit of electron microscopy. Phys. Rev. Lett. 1995, 74, 399–402.
ORDER REPRINTS

Characterization of Environmental Nanoparticles 2745

38. Chapman, J.N. The investigation of magnetic domain structures in thin foils by
electron microscopy. J. Phys. D: Appl. Phys. 1984, 17, 623–647.
39. Hiemenz, P.C.; Rajagopalan, R. Principles of Colloid and Surface Chemistry,
3rd Ed.; Marcel Dekker: New York, 1997; 566.
40. Park, K. Ph.D thesis, University of Minnesota, Minneapolis, 2003.
41. Delzeit, L.; Blake, D. A characterization of crystalline ice nanoclusters using
transmission electron microscopy. J. Geophys. Res. E: Planets 2001, 106 (E12),
33371–33379.
42. Sharma, R. Design and applications of environmental cell transmission electron
microscope for in situ observations of gas-solid reactions. Microsc. Microanal.
2001, 7 (6), 494–506.
43. Dragoo, A.L.; Robbins, C.R.; Hsu, S.M. A critical assessment of requirements
Downloaded by [Dalhousie University] at 06:38 20 August 2013

for ceramic powder characterization. Adv. Ceram. 1987, 21 (Ceram. Powder


Sci.), 711–720.
44. Maurice, P.A. Applications of atomic-force microscopy in environmental
colloid and surface chemistry. Colloid Surf. A-Physicochem. Eng. Asp. 1996,
107, 57–75.
45. Friedbacher, G.; Fuchs, H. Classification of scanning probe microscopies—
(technical report). Pure Appl. Chem. 1999, 71 (7), 1337–1357.
46. Sarid, D. Scanning Force Microscopy: with Applications to Electric, Magnetic
and Atomic Forces, Rev. Ed.; Oxford University Press: New York, 1994.
47. Nagy, K.L.; Cygan, R.T.; Hanchar, J.M.; Sturchio, N.C. Gibbsite growth
kinetics on gibbsite, kaolinite, and muscovite substrates: atomic force
microscopy evidence for epitaxy and an assessment of reactive surface area.
Geochim. Cosmochim. Acta 1999, 63 (16), 2337–2351.
48. Sutheimer, S.H.; Maurice, P.A.; Zhou, Q.H. Dissolution of well and poorly
crystallized kaolinites: Al speciation and effects of surface characteristics.
Am. Miner. 1999, 84 (4), 620–628.
49. Teng, H.H.; Dove, P.M.; DeYoreo, J.J. Reversed calcite morphologies induced
by microscopic growth kinetics: insight into biomineralization. Geochim.
Cosmochim. Acta 1999, 63 (17), 2507–2512.
50. Frisbie, C.D.; Rozsnyai, L.F.; Noy, A.; Wrighton, M.S.; Lieber, C.M.
Functional group imaging by chemical force microscopy. Science 1994,
265 (5181), 2071–2074.
51. Ito, S.; Yoshida, S.; Watanabe, T. Preparation of colloidal anatase TiO2
secondary submicroparticles by hydrothermal sol–gel method. Chem. Lett.
2000, (1), 70–71.
52. McKendry, R.; Theoclitou, M.E.; Abell, C.; Rayment, T. How much chemistry
is there in chemical force microscopy? Jpn. J. Appl. Phys. Part 1—Regul. Pap.
Short Notes Rev. Pap. 1999, 38 (6B), 3901–3907.
53. Proksch, R.B.; Schaffer, T.E.; Moskowitz, B.M.; Dahlberg, E.D.; Bazylinski,
D.A.; Frankel, R.B. Magnetic force microscopy of the submicron magnetic
assembly in a magnetotactic bacterium. Appl. Phys. Lett. 1995, 66 (19),
2582–2584.
54. Chen, X.; Davies, M.C.; Roberts, C.J.; Tendler, S.J.B.; Williams, P.M.;
Burnham, N.A. Optimizing phase imaging via dynamic force curves. Surf. Sci.
2000, 460 (1–3), 292–300.
ORDER REPRINTS

2746 Burleson, Driessen, and Penn

55. Wiesendanger, R. Scanning Probe Microscopy Analytical Methods; Springer:


Berlin, 1998.
56. Kirstein, S. Scanning near-field optical microscopy. Curr. Opin. Colloid
Interface Sci. 1999, 4 (4), 256–264.
57. Heinzelmann, H.; Pohl, D.W. Scanning near-field optical microscopy. Appl.
Phys. A 1994, A59 (2), 89–101.
58. Betzig, E.; Trautman, J.K. Near-field optics: microscopy, spectroscopy, and
surface modification beyond the diffraction limit. Science 1992, 257 (5067),
189–195.
59. Eggleston, C.M. Scanning Probe Microscopy of Clay Minerals; Clay Minerals
Society, 1994; Vol. 7.
60. Bickmore, B.R.; Hochella, M.F.; Bosbach, D.; Charlet, L. Methods for
Downloaded by [Dalhousie University] at 06:38 20 August 2013

performing atomic force microscopy imaging of clay minerals in aqueous


solutions. Clays Clay Miner. 1999, 47 (5), 573–581.
61. Goldstein, J.I.; Newbury, D.E.; Echlin, P.; Joy, D.C.; Romig, A.D., Jr.
Scanning Electron Microscopy and X-Ray Microanalysis, 2nd Ed.; Plenum
Press: New York, 1992.
62. Newbury, D.E.; Williams, D.B. The electron microscope: the materials
characterization tool of the millennium. Acta Metall. 2000, 48 (1), 323–346.
63. Scott, J.H.J. Analytical advances in the sem. Anal. Bioanal. Chem. 2003,
375 (1), 38–40.
64. Maynard, A.D. Overview of methods for analyzing single ultrafine particles.
Philos. Trans. R. Soc. London, Ser. A 2000, 358 (1775), 2593–2610.
65. Humphreys, F.J. Grain and subgrain characterization by electron backscatter
diffraction. J. Mater. Sci. 2001, 36 (16), 3833–3854.
66. Randle, V.; Caul, M. Representation of electron backscatter diffraction data.
Mater. Sci. Technol. 1996, 12 (10), 844–850.
67. Field, D.P. Recent advances in the application of orientation imaging.
Ultramicroscopy 1997, 67 (1–4), 1–9.
68. Dingley, D.J.; Randle, V. Microtexture determination by electron back-scatter
diffraction. J. Mater. Sci. 1992, 27 (17), 4545–4566.
69. Goetze, J. Potential of cathodoluminescence (cl) microscopy and spectroscopy
for the analysis of minerals and materials. Anal. Bioanal. Chem. 2002, 374 (4),
703–708.
70. Petruk, W. Imaging of minerals, ores and related products to determine mineral
characteristics. Miner. Metall. Process 2002, 19 (1), 50–56.
71. Strausser, Y.E. AES: Auger Electron Spectroscopy. In Encyclopedia of
Materials Characterization; Brundle, R.C., Evans, C.A., Wilson, S., Eds.;
Butterworth-Heinemann: Boston, 1992; 310–323.
72. Berube, K.A.; Jones, T.P.; Williamson, B.J.; Winters, C.; Morgan, A.J.;
Richards, R.J. Physicochemical characterization of diesel exhaust particles:
factors for assessing biological activity. Atmos. Environ. 1999, 33 (10),
1599–1614.
73. Ebert, M.; Inerle-Hof, M.; Weinbruch, S. Environmental scanning electron
microscopy as a new technique to determine the hygroscopic behaviour of
individual aerosol particles. Atmos. Environ. 2002, 36 (39–40), 5909–5916.
ORDER REPRINTS

Characterization of Environmental Nanoparticles 2747

74. Schleicher, B.; Tapper, U.; Kauppinen, E.I.; Martin, M.; Roschier, L.;
Paalanen, M.; Wernsdorfer, W.; Benoit, A. Magnetization reversal measure-
ments of size-selected iron oxide particles produced via an aerosol route. Appl.
Organomet. Chem. 1998, 12 (5), 315–320.
75. Weinbruch, S.; van Aken, P.; Ebert, M.; Thomassen, Y.; Skogstad, A.;
Chashchin, V.P.; Nikonov, A. The heterogeneous composition of working
place aerosols in a nickel refinery: a transmission and scanning electron
microscope study. J. Environ. Monit. 2002, 4 (3), 344–350.
76. Laskin, A.; Cowin, J.P. Automated single-particle SEM/EDX analysis of
submicrometer particles down to 0.1 mm. Anal. Chem. 2001, 73 (5), 1023–1029.
77. Kasparian, J.; Frejafon, E.; Rambaldi, P.; Yu, J.; Vezin, B.; Wolf, J.P.; Ritter,
P.; Viscardi, P. Characterization of urban aerosols using sem-microscopy,
Downloaded by [Dalhousie University] at 06:38 20 August 2013

X-ray analysis and lidar measurements. Atmos. Environ. 1998, 32 (17),


2957–2967.
78. Swietlicki, E.; Martinsson, B.G.; Kristiansson, P. The use of PIXE and
complementary ion beam analytical techniques for studies of atmospheric
aerosols. Nucl. Instrum. Methods Phys. Res., Sect. B 1996, 109/110, 385–394.
79. Johansson, S.A.E.; Campbell, J.L.; Malmqvist, K.G. Particle-Induced X-Ray
Emission Spectroscopy (Pixe). In Chemical Analysis; Winefordner, J.D., Ed.;
John Wiley & Sons, Inc.: New York, 1995; Vol. 133.
80. Miranda, J. Studies of atmospheric aerosols in large urban areas using pixe: an
overview. Nucl. Instrum. Methods Phys. Res., Sect. B 1996, 109/110, 439–444.
81. Maenhaut, W.; Cafmeyer, J. Long-term atmospheric aerosol study at urban
and rural sites in belgium using multi-elemental analysis by particle-enduced
X-ray emission spectrometry and short-irradiation instrumental neutron
activation analysis. X-Ray Spectrom. 1998, 27 (4), 236–246.
82. Malmqvist, K.G. Particle-induced X-ray emission—a quantitative technique
suitable for microanalysis. Mikrochim. Act. Suppl. 1996, 13 (Microbeam and
Nanobeam Analysis), 117–133.
83. Malmqvist, K. The nuclear microprobe—advances in micro PIXE and
complementary techniques. X-Ray Spectrom. 1995, 24 (5), 226–234.
84. Campbell, J.L.; Higuchi, D.; Maxwell, J.A.; Teesdale, W.J. Quantitative PIXE
microanalysis of thick specimens. Nucl. Instrum. Methods Phys. Res., Sect. B
1993, B77 (1–4), 95–109.
85. Maura de Miranda, R.; Andrade, M.D.F.; Worobiec, A.; van Grieken, R.
Characterisation of aerosol particles in the Sao Paulo metropolitan area.
Atmos. Environ. 2002, 36 (2), 345–352.
86. Salma, I.; Maenhaut, W.; Zaray, G. Comparative study of elemental mass size
distributions in urban atmospheric aerosol. J. Aerosol Sci. 2001, 33 (2),
339–356.
87. Eleftheriadis, K.; Colbeck, I. Coarse atmospheric aerosol: size distributions of
trace elements. Atmos. Environ. 2001, 35 (31), 5321–5330.
88. Briggs, D.; Seah, M.P. Practical Surface Analysis, 2nd Ed.; John Wiley & Sons:
Chichester, 1992; Vol. 1—Auger and X-ray Photoelectron Spectroscopy.
89. Anschutz, A.J.; Penn, R.L. Reduction of crystalline iron(III) oxyhydroxides
using hydroquinone: influence of phase and particle size, in prep.
ORDER REPRINTS

2748 Burleson, Driessen, and Penn

90. Gunther, S.; Kaulich, B.; Gregoratti, L.; Kiskinova, M. Photoelectron


microscopy and applications in surface and materials science. Progress in
Surface Science 2002, 70 (4–8), 187–260.
91. Hoenig, M. Preparation steps in environmental trace element analysis-facts
and traps. Talanta 2001, 54, 1021–1038.
92. Ford, R.G. Rates of hydrous ferric oxide crystallization and the influence on
coprecipitated arsenate. Environ. Sci. Technol. 2002, 36, 2459–2463.
93. Ford, R.G.; Kemner, K.M.; Bertsch, P.M. Influence of sorbate-sorbent
interactions on the crystallization kinetics of nickel- and lead-ferrihydrite
coprecipitates. Geochim. Cosmochim. Acta 1999, 63 (1), 39–48.
94. Cornell, R.M.; Schneider, W. Formation of goethite from ferrihydrite at
physiological pH under the influence of cysteine. Polyhedron 1989, 8 (2),
Downloaded by [Dalhousie University] at 06:38 20 August 2013

149–155.
95. Van Loon, J.C. Analytical Atomic Absorption Spectroscopy: Selected Methods;
Academic Press: New York, 1980.
96. Welz, B.; Sperling, M. Atomic Absorption Spectrometry, 3rd Ed.; Wiley-VCH:
Weinheim, 1999.
97. Balaram, V. Recent trends in the inorganic analysis of environmental
materials. Res. J. Chem. Environ. 2002, 6 (2), 69–79.
98. Taylor, V.F.; Toms, A.; Longerich, H.P. Acid digestion of geological and
environmental samples using open-vessel focused microwave digestion. Anal.
Bioanal. Chem. 2002, 372, 360–365.
99. Ridley, I.W. The ICP-MS laser microprobe: a new geochemical tool. Trends in
Geochemistry 2000, 1, 1–14.
100. Hill, S.J.; Arowolo, T.A.; Butler, O.T.; Cook, J.M.; Cresser, M.S.;
Harrington, C.; Miles, D.L. Atomic spectrometry update. Environmental
analysis. J. Anal. At. Spectrom. 2003, 18, 170–202.
101. Kurfurst, U. Solid Sample Analysis; Springer-Verlag: Berlin Heidelberg, 1998.
102. Dean, J.R. Atomic Absorption and Plasma Spectroscopy, 2nd Ed.; John Wiley
& Sons: Chichester, 1997.
103. Ebdon, L.; Evans, E.H.; Fisher, A.; Hill, S.J. An Introduction to Analytical
Atomic Spectrometry; John Wiley & Sons: Chichester, 1998.
104. Jarvis, K.E.; Gray, A.L.; Houk, R.S. Handbook of Inductively Coupled Plasma
Mass Spectrometry; Blackie: Glasgow & London, 1992.
105. Gomez, D.; Smichowski, P.; Polla, G.; Ledesma, A.; Resnizky, S.; Rosa, S.
Fractionation of elements by particle size of ashes ejected from copahue
volcano, argentina. J. Environ. Monit. 2002, 4, 972–977.
106. Skoog, D.A.; Holler, F.J.; Nieman, T.A. Principles of Instrumental Analysis,
5th Ed.; Brooks/Cole, 1997.
107. Cullity, B.D.; Stock, S.R. Elements of X-Ray Diffraction, 3rd Ed.; Prentice
Hall: Upper Saddle River, 2001.
108. Toney, M.F. XRD: X-ray diffraction. In Encyclopedia of Materials
Characterization; Brundle, R.C., Evans, C.A., Wilson, S., Eds.; Butterworth-
Heinemann: Boston, 1992; 198–213.
109. Huber, C. X-ray diffraction characterization of nanophase materials. In
Handbook of Nanophase Materials; Goldstein, A.N., Ed.; Marcel Dekker:
New York, 1997; 317–336.
ORDER REPRINTS

Characterization of Environmental Nanoparticles 2749

110. Tarassov, M.; Mihailova, B.; Tarassova, E.; Konstantinov, L. Chemical


composition and vibrational spectra of tungsten-bearing goethite and hematite
from western rhodopes, Bulgaria. Eur. J. Mineral. 2002, 14, 977–986.
111. Matyi, R.J.; Schwart, L.H.; Butt, J.B. Partice size, particle size distribution,
and related measurements of supported metal catalysts. Cat. Rev. -Sci. Eng.
1987, 29, 41–99.
112. Zanchet, D.; Hall, B.D.; Ugarte, D. X-ray characterization of nanoparticles.
In Characterization of Nanophase Materials; Wang, Z.L., Ed.; Wiley-VCH:
Weinheim, New York, 2000; 13–36.
113. Davis, B.L.; Jenkins, R.; McCarthy, G.J.; Smith, D.K.; Wong-Ng, W.
Specimen preparation in X-ray diffraction. In A Practical Guide for the
Preparation of Specimens for X-Ray Fluorescence and X-Ray Diffraction
Downloaded by [Dalhousie University] at 06:38 20 August 2013

Analysis; Buhrke, V.E., Jenkins, R., Smith, D.K., Eds.; John Wiley & Sons:
New York, 1998; 123–170.
114. Blanton, T.N.; Hamilton, R.F.; Iyenger, S.S.; Jenkins, R.; McMurdie, H.F.;
Nelson, D.E.; Rasmussen, S.E.; Rendle, D.F.; Roach, G.I.D.; Ryba, E.R.;
Smith, D.K.; Wong-Ng, W. Specific areas of specimen preparation in X-ray
powder diffraction. In A Practical Guide for the Preparation of Specimens for
X-Ray Fluorescence and X-Ray Diffraction Analysis; Buhrke, V.E., Jenkins,
R., Smith, D.K., Eds.; John Wiley & Sons: New York, 1998; 171–216.
115. Antonio, M.R. EXAFS: Extended X-ray absorption fine structure. In
Encyclopedia of Materials Characterization; Brundle, R.C., Evans, C.A.,
Wilson, S., Eds.; Butterworth-Heinemann: Boston, 1992; 214–226.
116. Fendorf, S.E.; Sparks, D.L.; Lamble, G.M.; Kelley, M.J. Applications of
X-ray adsorption fine-structure spectroscopy to soils. Soil Sci. Soc. Am. J.
1994, 58, 1583–1595.
117. Waychunas, G.A. Structure, aggregation and characterization of nanoparti-
cles. In Nanoparticles and the Environment; Banfield, J.F., Navrotsky, A., Eds.;
Mineralogical Society of America: Washington, D.C., 2001; Vol. 44, 105–166.
118. Waychunas, G.A.; Fuller, C.C.; Davis, J.A. Surface complexation and
precipitate geometry for aqueous Zn(II) sorption on ferrihydrite I: X-ray
absorption extended fine structure spectroscopy analysis. Geochim.
Cosmochim. Acta 2002, 66 (7), 1119–1137.
119. Waychunas, G.A.; Fuller, C.C.; Rea, B.A.; Davis, J.A. Wide angle X-ray
scattering (WAXS) study of ‘‘two-line’’ ferrihydrite structure: effect of
arsenate sorption and counterion variation and comparison with EXAFE
results. Geochim. Cosmochim. Acta 1996, 60 (10), 1765–1781.
120. Cheng, L.; Sturchio, N.C.; Bedzyk, M.J. Local structure of Co2þ incorporated
at the calcite surface. An X-ray standing wave and sexafs study. Phys. Rev. B:
Condens. Matter 2000, 61 (7), 4877–4883.
121. Lagarde, P. Surface X-ray absorption spectroscopy principles and some
examples of applications. Ultramicroscopy 2001, 86 (3/4), 255–263.
122. Norman, D. SEXAFS/NEXAFS: Surface Extended X-Ray Absorption Fine
Structure and Near Edge X-Ray Absorption Fine Structure. In Encyclopedia
of Materials Characterization; Brundle, R.C., Evans, C.A., Wilson, S., Eds.;
Butterworth-Heinemann: Boston, 1992; 227–239.
ORDER REPRINTS

2750 Burleson, Driessen, and Penn

123. Stor, J. SEXAFS: Everything You Always Wanted to Know about SEXAFS
but Were Afraid to Ask. In X-Ray Absorption: Principles, Applications
Techniques of EXAFS, SEXAFS and XANES; Koningsberger, D.C., Prins, R.,
Eds.; John Wiley & Sons: New York, 1988; Vol. 92, 443–572.
124. Ingle, J., James D.; Crouch, S. Spectrochemical Analysis; Prentice-Hall: Upper
Saddle River, NJ, 1988.
125. Schwertmann, U.; Friedl, J.; Stanjek, H.; Murad, E.; Bender Koch, C. Iron
oxides and smectites in sediments from the Atlantis II deep, red sea. Eur. J.
Mineral. 1998, 10 (5), 953–967.
126. Woolridge, M.S.; Danczyk, S.A.; Wu, J. Demonstration of gas-phase
combustion synthesis of nanosized particles using a hybrid burner.
Downloaded by [Dalhousie University] at 06:38 20 August 2013

Nanostruct. Mater. 1999, 11 (7), 955–964.


127. Lyon, L.A.; Keating, C.D.; Fox, A.P.; Baker, B.E.; He, L.; Nicewarner, S.R.;
Mulvaney, S.P.; Natan, M.J. Raman spectroscopy. Anal. Chem. 1998, 70,
341R–361R.
128. Pelletier, M.J. Analytical Applications of Raman Spectroscopy; Blackwell
Science: Oxford, 1999.
129. Weber, W.H.; Merlin, R. Raman Scattering in Materials Science; Springer-
Verlag: Berlin, 2000.
130. Mulvaney, S.P.; Keating, C.D. Raman spectroscopy. Anal. Chem. 2000, 72,
145R–157R.
131. de Faria, D.L.A.; Silva, S.V.; de Oliveira, M.T. Raman microspectroscopy of
some iron oxides and oxyhydroxides. J. Raman Spectrosc. 1997, 28, 873–878.
132. Driessen, M.D.; Goodman, A.L.; Miller, T.M.; Zaharias, G.A.; Grassian,
V.H. Gas-phase photooxidation of trichloroethylene on TiO2 and ZnO:
influence of trichloroethylene pressure, oxygen pressure, and the photocatalyst
surface on the product distribution. J. Phys. Chem. B 1998, 102 (3), 549–556.
133. Driessen, M.D.; Grassian, V.H. Oxidation of methyl to methoxy group on
oxidized Cu/SiO2. J. Phys. Chem. 1995, 99 (45), 16519–16522.
134. Driessen, M.D.; Grassian, V.H. A comprehensive study of the reactions of
methyl fragments from methyl iodide dissociation on reduced and oxidized
silica-supported copper nanoparticles. J. Am. Chem. Soc. 1997, 119 (7),
1697–1707.
135. Khaleel, A.; Dellinger, B. FTIR investigation of adsorption and chemical
decomposition of CCl4 by high surface-area aluminum oxide. Environ. Sci.
Technol. 2002, 36 (7), 1620–1624.
136. Rancourt, D.G. Magnetism of Earth, Planetary, and Environmental
Nanomaterials. In Nanoparticles and the Environment; Banfield, J.F.,
Navrotsky, A., Eds.; Mineralogical Society of America: Washington, D.C.,
2001; Vol. 44, 217–292.
137. Sorenson, C.M. Magnetism. In Nanoscale Materials in Chemistry; Klabunde,
K.J., Ed.; John Wiley & Sons, Inc.: New York, 2001; 169–221.
138. Mulay, L.N.; Mulay, I.L. Magnetometry: instrumentation and analytical
applications including catalysis, bioscience, geoscience, and amorphous
materials. Anal. Chem. 1982, 54, 216R–227R.
ORDER REPRINTS

Characterization of Environmental Nanoparticles 2751

139. Kopcewicz, B.; Kopcewicz, M. Mossbauer study of the structure of iron-


containing atmospheric aerosols. In Analytical Chemistry of Aerosols; Spurny,
K.R., Ed.; Lewis Publishers, CRC, Press LLC: Boca Raton, 1999; 185–196.
140. Stevens, J.G.; Bowen, L.H. Mossbauer spectroscopy. Anal. Chem. 1984, 56,
199R–212R.
141. Dollimore, D.; Spooner, P.; Turner, A. The BET method of analysis of gas
adsorption data and its relevance to the calculation of surface areas. Surf.
Technol. 1976, 4 (2), 121–160.
142. Brunauer, S.; Emmett, P.H.; Teller, E. Adsorption of gases in multimolecular
layers. J. Am. Chem. Soc. 1938, 60, 309–319.
143. Toledo-Antonio, J.A.; Gutierrez-Baez, R.; Sebastian, P.J.; Vazquez, A.
Thermal stability and structural deformation of rutile SnO2 nanoparticles.
Downloaded by [Dalhousie University] at 06:38 20 August 2013

J. Solid State Chem. 2003, 174 (2), 241–248.


144. Yen, F.S.; Chen, W.C.; Yang, J.M.; Hong, C.T. Crystallite size variations of
nanosized Fe2O3 powders during g- to a-phase transformation. Nano Lett.
2002, 2 (3), 245–252.
145. Mogyorosi, K.; Dekany, I.; Fendler, J.H. Preparation and characterization
of clay mineral intercalated titanium dioxide nanoparticles. Langmuir 2003,
19 (7), 2938–2946.
146. Muller, R.H.; Mehnert, W. Particle and Surface Characterisation Methods;
Medpharm Scientific Publishers: Stuttgart, 1997.
147. De Jaeger, N.; Demeyere, H.; Finsy, R.; Sneyers, R.; Vanderdeelen, J.; Van
der Meeren, P.; Van Laethem, M. Particle sizing by photon correlation
spectroscopy. Part I. Monodisperse latices: influence of scattering angle and
concentration of dispersed material. Part. Part. Syst. Char. 1991, 8 (3),
179–186.
148. Finsy, R.; De Jaeger, N. Particle sizing by photon correlation spectroscopy.
Part II. Average values. Part. Part. Syst. Char. 1991, 8 (3), 187–193.
149. Finsy, R.; De Jaeger, N.; Sneyers, R.; Gelade, E. Particle sizing by photon
correlation spectroscopy. Part III. Mono and bimodal distributions and data
analysis. Part. Part. Syst. Char. 1992, 9 (2), 125–137.
150. Finsy, R. Particle sizing by quasi-elastic light scattering. Adv. Colloid
Interface Sci. 1994, 52, 79–143.
151. Chu, B.; Liu, T. Characterization of nanoparticles by scattering techniques.
J. Nanopart. Res. 2000, 2 (1), 29–41.
152. Cecere, D.; Bruno, A.; Minutolo, P.; D’Alessio, A. DLS measurements on
nanoparticles produced in laminar premixed flames. Synth. Met. 2003, 139 (3),
653–656.
153. Sipos, P.; Berkesi, O.; Tombacz, E.; St. Pierre, T.G.; Webb, J. Formation of
spherical iron(III) oxyhydroxide nanoparticles sterically stabilized by chitosan
in aqueous solutions. J. Inorg. Biochem. 2003, 95 (1), 55–63.
154. Jeevanandam, P.; Klabunde, K.J. Redispersion and reactivity studies on
surfactant-coated magnesium oxide nanoparticles. Langmuir 2003, 19 (13),
5491–5495.
155. Fergenson, D.P.; Song, X.-H.; Ramadan, Z.; Allen, J.O.; Hughes, L.S.; Cass,
G.R.; Hopke, P.K.; Prather, K.A. Quantification of ATOFMS data by
multivariate methods. Anal. Chem. 2001, 73 (15), 3535–3541.
ORDER REPRINTS

2752 Burleson, Driessen, and Penn

156. Gard, E.; Mayer, J.E.; Morrical, B.D.; Dienes, T.; Fergenson, D.P.; Prather,
K.A. Real-time analysis of individual atmospheric aerosol particles: design
and performance of a portable atofms. Anal. Chem. 1997, 69 (20), 4083–4091.
157. Hughes, L.S.; Allen, J.O.; Kleeman, M.J.; Johnson, R.J.; Cass, G.R.; Gross,
D.S.; Gard, E.E.; Gaelli, M.E.; Morrical, B.D.; Fergenson, D.P.; Dienes, T.;
Noble, C.A.; Liu, D.-Y.; Silva, P.J.; Prather, K.A. Size and composition
distribution of atmospheric particles in southern California. Environ. Sci.
Technol. 1999, 33 (20), 3506–3515.
158. Galli, M.; Guazzotti, S.A.; Prather, K.A. Improved lower particle size limit for
aerosol time-of-flight mass spectrometry. Aerosol Sci. Technol. 2001, 34 (4),
381–385.
159. Gard, E.E.; Kleeman, M.J.; Gross, D.S.; Hughes, L.S.; Allen, J.O.; Morrical,
Downloaded by [Dalhousie University] at 06:38 20 August 2013

B.D.; Fergenson, D.P.; Dienes, T.; Galli, M.E.; Johnson, R.J.; Cass, G.R.;
Prather, K.A. Direct observation of heterogeneous chemistry in the atmo-
sphere. Science 1998, 279 (5354), 1184–1187.
160. Gross, D.S.; Warren, B.S.; Barron, A.R.; Liu, D.-Y.; Wenzel, R.J.; Prather,
K.A. Book of Abstracts, 219th ACS National Meeting, San Francisco, CA,
March 26–30, 2000; CHED-237.
161. Liu, D.-Y.; Wenzel, R.J.; Prather, K.A. Aerosol time-of-flight mass spectrom-
etry during the atlanta supersite experiment: 1. Measurements. J. Geophys.
Res. D: Atmos. 2003, 108 (D7), SOS 14/1–SOS 14/16.
162. Middlebrook, A.M.; Murphy, D.M.; Lee, S.-H.; Thomson, D.S.; Prather,
K.A.; Wenzel, R.J.; Liu, D.-Y.; Phares, D.J.; Rhoads, K.P.; Wexler, A.S.;
Johnston, M.V.; Jimenez, J.L.; Jayne, J.T.; Worsnop, D.R.; Yourshaw, I.;
Seinfeld, J.H.; Flagan, R.C. A comparison of particle mass spectrometers
during the 1999 Atlanta supersite project. J. Geophys. Res. D: Atmos. 2003,
108 (D7), SOS 12/1–SOS 12/13.
163. Wenzel, R.J.; Liu, D.-Y.; Edgerton, E.S.; Prather, K.A. Aerosol time-of-flight
mass spectrometry during the Atlanta supersite experiment: 2. Scaling
procedures. J. Geophys. Res. D: Atmos. 2003, 108 (D7), SOS 15/1–SOS 15/8.
164. Knutson, E.O.; Whitby, K.T. Aerosol classification by electric mobility:
apparatus, theory, and applications. J. Aerosol Sci. 1975, 6 (6), 443–451.
165. Hinds, W.C. Aerosol Technology: Properties, Behavior, and Measurement of
Airborne Particles, 2nd Ed.; John Wiley & Sons, Inc.: New York, 1999.
166. Friedlander, S.K. Smoke, Dust, and Haze, 2nd Ed.; Oxford University Press:
New York, 2000.
167. Weber, A.P.; Baltensperger, U.; Gaeggeler, H.W.; Schmidt-Ott, A. In situ
characterization and structure modification of agglomerated aerosol particles.
J. Aerosol Sci. 1996, 27 (6), 915–929.
168. Tobias, H.J.; Beving, D.E.; Ziemann, P.J.; Sakurai, H.; Zuk, M.; McMurry,
P.H.; Zarling, D.; Waytulonis, R.; Kittelson, D.B. Chemical analysis of diesel
engine nanoparticles using a nano-dma/thermal desorption particle beam
mass spectrometer. Environ. Sci. Technol. 2001, 35 (11), 2233–2243.
169. Higgins, K.J.; Jung, H.; Kittelson, D.B.; Roberts, J.T.; Zachariah, M.R.
Kinetics of diesel nanoparticle oxidation. Environ. Sci. Technol. 2003, 37 (9),
1949–1954.
ORDER REPRINTS

Characterization of Environmental Nanoparticles 2753

170. Franz, B.; Eckhardt, T.; Kauffeldt, T.; Roth, P. H2O2 addition to diesel engine
exhaust gas and its effect on particles. J. Aerosol Sci. 2000, 31 (4), 415–426.
171. McMurry, P.H.; Litchy, M.; Huang, P.-F.; Cai, X.; Turpin, B.J.; Dick, W.D.;
Hanson, A. Elemental composition and morphology of individual particles
separated by size and hygroscopicity with the TDMA. Atmos. Environ. 1996,
30 (1), 101–108.
172. Rader, D.J.; McMurry, P.H. Application of the tandem differential mobility
analyzer to studies of droplet growth or evaporation. J. Aerosol Sci. 1986,
17 (5), 771–787.
173. Higgins, K.J.; Jung, H.; Kittelson, D.B.; Roberts, J.T.; Zachariah, M.R. Size-
selected nanoparticle chemistry: kinetics of soot oxidation. J. Phys. Chem. A
Downloaded by [Dalhousie University] at 06:38 20 August 2013

2002, 106 (1), 96–103.


Request Permission or Order Reprints Instantly!

Interested in copying and sharing this article? In most cases, U.S. Copyright
Law requires that you get permission from the article’s rightsholder before
using copyrighted content.

All information and materials found in this article, including but not limited
to text, trademarks, patents, logos, graphics and images (the "Materials"), are
the copyrighted works and other forms of intellectual property of Marcel
Dekker, Inc., or its licensors. All rights not expressly granted are reserved.
Downloaded by [Dalhousie University] at 06:38 20 August 2013

Get permission to lawfully reproduce and distribute the Materials or order


reprints quickly and painlessly. Simply click on the "Request Permission/
Order Reprints" link below and follow the instructions. Visit the
U.S. Copyright Office for information on Fair Use limitations of U.S.
copyright law. Please refer to The Association of American Publishers’
(AAP) website for guidelines on Fair Use in the Classroom.

The Materials are for your personal use only and cannot be reformatted,
reposted, resold or distributed by electronic means or otherwise without
permission from Marcel Dekker, Inc. Marcel Dekker, Inc. grants you the
limited right to display the Materials only on your personal computer or
personal wireless device, and to copy and download single copies of such
Materials provided that any copyright, trademark or other notice appearing
on such Materials is also retained by, displayed, copied or downloaded as
part of the Materials and is not removed or obscured, and provided you do
not edit, modify, alter or enhance the Materials. Please refer to our Website
User Agreement for more details.

Request Permission/Order Reprints

Reprints of this article can also be ordered at


http://www.dekker.com/servlet/product/DOI/101081ESE200027029

Das könnte Ihnen auch gefallen