ArticlePDF AvailableLiterature Review

Nanoparticle analysis and characterization methodology in environmental risk assessment of engineered nanoparticles

Authors:

Abstract and Figures

Environmental risk assessments of engineered nanoparticles require thorough characterization of nanoparticles and their aggregates. Furthermore, quantitative analytical methods are required to determine environmental concentrations and enable both effect and exposure assessments. Many methods still need optimization and development, especially for new types of nanoparticles in water, but extensive experience can be gained from the fields of environmental chemistry of natural nanomaterials and from fundamental colloid chemistry. This review briefly describes most methods that are being exploited in nanoecotoxicology for analysis and characterization of nanomaterials. Methodological aspects are discussed in relation to the fields of nanometrology, particle size analysis and analytical chemistry. Differences in both the type of size measures (length, radius, aspect ratio, etc.), and the type of average or distributions afforded by the specific measures are compared. The strengths of single particle methods, such as electron microscopy and atomic force microscopy, with respect to imaging, shape determinations and application to particle process studies are discussed, together with their limitations in terms of counting statistics and sample preparation. Methods based on the measurement of particle populations are discussed in terms of their quantitative analyses, but the necessity of knowing their limitations in size range and concentration range is also considered. The advantage of combining complementary methods is highlighted.
Content may be subject to copyright.
Nanoparticle analysis and characterization methodologies
in environmental risk assessment of engineered nanoparticles
Martin Hassello
¨
v Æ James W. Readman Æ
James F. Ranville Æ Karen Tiede
Accepted: 28 April 2008 / Published online: 16 May 2008
Ó Springer Science+Business Media, LLC 2008
Abstract Environmental risk assessments of engineered
nanoparticles require thorough characterization of nano-
particles and their aggregates. Furthermore, quantitative
analytical methods are required to determine environmen-
tal concentrations and enable both effect and exposure
assessments. Many methods still need optimization and
development, especially for new types of nanoparticles in
water, but extensive experience can be gained from the
fields of environmental chemistry of natural nanomaterials
and from fundamental colloid chemistry. This review
briefly describes most methods that are being exploited in
nanoecotoxicology for analysis and characterization of
nanomaterials. Methodological aspects are discussed in
relation to the fields of nanometrology, particle size anal-
ysis and analytical chemistry. Differences in both the type
of size measures (length, radius, aspect ratio, etc.), and the
type of average or distributions afforded by the specific
measures are compared. The strengths of single particle
methods, such as electron microscopy and atomic force
microscopy, with respect to imaging, shape determinations
and application to particle process studies are discussed,
together with their limitations in terms of counting statis-
tics and sample preparation. Methods based on the
measurement of particle populations are discussed in terms
of their quantitative analyses, but the necessity of knowing
their limitations in size range and concentration range is
also considered. The advantage of combining comple-
mentary methods is highlighted.
Keywords Nanoparticles Nanoaggregates
Nanometrology Analytical chemistry
Particle size analysis
Introduction
Due to the extensive current, and foreseen future invest-
ments, in nanotechnology, nanoparticles used in consumer
products, industrial applications and health care technology
are likely to enter the environment (Aitken et al. 2006;
Roco 2005). To ensure sustainable development of nano-
technology, there is a need for risk assessments of
engineered nanoparticles (ENP) introduced from various
applications (Colvin 2003; Maynard et al. 2006). Such risk
assessments, require proper tools and methodologies to
carry out both effect and exposure assessments (EPA 2007;
Maynard et al. 2006; SCENIHR 2005; Crane and Handy
2007). Conventionally, exposure assessment is recom-
mended to include both a modeling and a measurement
approach (Holt et al. 2000); both approaches require
instrumentation and analytical methods. Prediction of
environmental concentrations of ENP through modeling is
based on emission scenarios (from production volumes and
M. Hassello
¨
v(&)
Department of Chemistry, University of Gothenburg,
Gothenburg 41296, Sweden
e-mail: martin.hassellov@chem.gu.se
J. W. Readman
Plymouth Marine Laboratory, Prospect Place,
Plymouth PL1 3DH, UK
J. F. Ranville
Department of Chemistry & Geochemistry, Colorado School of
Mines, Golden, CO 80401, USA
K. Tiede
Central Science Laboratory, Sand Hutton, York YO41 1LZ, UK
K. Tiede
Environment Department, University of York, Heslington,
York YO10 5DD, UK
123
Ecotoxicology (2008) 17:344–361
DOI 10.1007/s10646-008-0225-x
life cycle assessments) and partitioning parameters (fate
and behavior). Presently, little is known about the fate and
behavior parameters of ENP. Hence, development of suit-
able analytical methods are required to determine
concentrations and nanoparticle characteristics in complex
environmental matrices such as water, soil, sediment,
sewage sludge and biological specimens. The approach for
prediction of environmental concentrations through mod-
eling requires validation through measurement of actual
environmental concentrations. For ENPs that are only
recently being introduced into the environment, extremely
sensitive methods are required. Although direct observa-
tions are not hampered by the underlying assumptions of
exposure modeling, it is very important to assure that direct
observations are representative in time and space for the
regional setting to which the observation will be allocated
(local or regional).
ENP differ from most conventional ‘dissolved’ chem-
icals in terms of their heterogeneous distributions in size,
shape, surface charge, composition, degree of dispersion,
etc. Therefore, it is not only important to determine their
concentrations, but also several other metrics.
In addition to exposure assessment requirements, it is
essential that characterization of ENP dispersion states (i.e.
aggregated or dispersed), and measurements of ‘steady-
state’ concentrations are used in effect assessment test
systems (e.g., toxicity testing). It has been found that the
ENP concentrations are often not sustained in dispersions
throughout an experiment (Federici et al. 2007). Although
this need was not recognized in the pioneer studies in
nanoecotoxicology, it is now starting to be implemented in
most effects experiments. In a recent review (Hansen et al.
2007), it was shown that although size determinations are
becoming more common (17–96% of exposure and effects
studies), other relevant characterization properties are
rarely determined (e.g., surface area in only 6–33% of
studies). An additional complication relates to stability. For
example, Fig. 1 demonstrates that Buckminster fullerenes
readily degrade and are highly reactive (Taylor et al.
1991). Indeed, it is the reactions of Buckminster fullerenes
that render them of particular interest when investigating
their potential applications in nanotechnology (Taylor
2006). This reactivity has substantial implications in
interpretation of environmental behaviour and ecotoxico-
logical impact.
Assessing uptake and bioaccumulation in biological
matrices are essential and will be equally as challenging as
analyses of complex environmental media. Furthermore,
some good laboratory practices and harmonized methods
still need to be developed. Due to both the complexity of
the behavior of nanomaterials in dispersions and the
requirements for expertise in state-of-the-art methods in
ecotoxicology testing and nanoparticle characterization, the
necessity for interdisciplinary collaboration has been
highlighted (Crane and Handy 2007; Handy et al. 2008).
This paper focuses on mature and validated methods that
are commercially available and/or fairly easy to setup.
Consequently, highly specialized methods in the develop-
ment phase, or methods requiring large-scale facilities such
as synchrotron sources, are not discussed.
Nanometrology, analytical chemistry and particle
size analysis
The physical properties of nanoparticles are referred to as
metrics (Table 1) and the field of science which aims to
standardize physical measurements at the nanometer scale,
is called nanometrology. Even though nanometrology is a
young field regarding definitions and terminology, many
concepts are borrowed and adopted from the fields of
particle size analysis (Barth and Flippen 1995) and physi-
cal chemistry. In addition to the physical properties,
nanoparticles can be described by their chemical compo-
sition where the compound or species determined is called
the analyte (Table 1).
The metric ‘particle diameter is probably the most
commonly used descriptor of particle size, but a single
diameter value is only enough to describe a perfect
spherical particle. Non-spherical nanoparticles (or colloids)
are, however, common in the environment and it is actually
common for nanoparticles to have very large aspect ratios
(e.g., clay platelets, rods or fibrils). Many engineered
nanoparticles share these features (e.g., carbon nanotubes,
nanowires, nanoclays, nanorods). It has been shown that
the toxicity can be shape dependent (Pal et al. 2007), and
Fig. 1 C60 fullerene solutions (in toluene) stored under dark and
light conditions (Photograph courtesy of P. Frickers and J.W.
Readman). The photo illustrates the potential of using spectroscopic
methods to study these photochemical changes in structure or surface
chemistry
Nanoparticle analysis and characterization methods 345
123
nanoparticle reactivity can be dependent both on size and
shape (Madden and Hochella 2005). There are several
different diameter measures that correspond to an equiva-
lent size of a specific type (Table 2). Different particle size
analysis methods also yield different equivalent sizes
(Table 2), which is important to consider when comparing
size values obtained using different methods. Another
important feature in method comparisons is that different
techniques give different size averages, depending on if
they fundamentally rely on an instrument response to:
particle numbers, volume, mass or optical property (e.g.,
light scattering) (Table 3). These averages can be the same
for spherical, monodisperse particles (with an infinitely
narrow size distribution) This, however, is usually not the
Table 1 A list of physical
properties (metrics), and a list
chemical compositions, analytes
and respective associated
methods and instruments
a
For abbreviations see text
Instruments and methods
a
Physical properties/metrics
Diameter EM, AFM, Flow-FFF, DLS,
Volume Sed-FFF
Area EM, AFM
Mass LC-ESMS
Surface charge z-Potential, electrophoretic mobility
Crystal structure XRD, TEM-XRD (SAED)
Aspect ratio or other shape factor
Chemical composition/analytes
Elemental composition Bulk: ICP-MS, ICP-OES, single nanoparticle:
TEM-EDX, particle population: FFF-ICP-MS
Fluorophores Fluorescence spectroscopy
Fullerene (‘‘molecules’’) UV–vis, IR, NMR, MS, HPLC
Total organic carbon High temp chemical oxidation
Other properties not falling within the above classes
Aggregation state DLS, AFM, ESEM, etc.
Hydrophobicity Liquid–liquid extraction chromatography
Dissolution rate Dialysis or voltammetry or spectrometry
Surface chemistry, coating composition,
# of proton exchanging surface sites
Optical or X-ray spectroscopic methods,
acid–base titrations
Table 2 Different equivalent
sizes measured by different
methods
Equivalent spherical
size measures
Applies to method
Hydrodynamic diameter Flow-FFF, DLS Calculated from the measured
diffusion coefficient, using
Stokes–Einstein equation
Equivalent spherical volume
diameter
Sed-FFF (if known density),
LIBD, electrozone sensing
Buoyant mass Sed-FFF SedFFF µ Dd*V
Equivalent spherical mass
diameter
MS Assume a certain structure
Projected area Microscopy
Equivalent molar mass Ultrafiltration Molecular weight cutoff (MWCO),
defined from retention of proteins
Equivalent poresize diameter Particle filtration Filter poresize often defined as
maximum size that penetrates filter
Root mean square radius of
gyration
SLS mean square distances from center
of mass of point masses within the
particle
Aspect ratio Microscopy, combination
of light scattering
methods or different
FFF methods
The longest dimension divided by the
shortest for symmetrical particles
(e.g., rods & ellipsoids)
346 M. Hassello
¨
v et al.
123
case. Each method also has its limitations in applicable size
and concentration ranges (Table 4). Therefore, it has to be
taken into account that there may be part of the nanopar-
ticle (or nanoparticle-aggregate) size distribution that is
‘hidden’ for the applied method. Some relevant terms and
definitions from analytic chemistry, nanometrology and
particle size analysis is given in Table 5.
There are some special challenges for studies of ENPs in
environmental samples. The first challenge is that for envi-
ronmentally relevant concentrations (ng l
-1
–pg l
-1
), the
detection limits for most methods are not sufficiently low.
The second challenge is that in environmental samples there
is a high background of natural and unintentionally produced
nanoparticles (Banfield and Navrotsky 2001; Filella 2007;
Hochella and Madden 2005; Lead and Wilkinson 2006;
Waychunas et al. 2005; Wigginton et al. 2007).
A strategy for coping with these challenges may be to
combine existing and new methods that afford both a
screening capability and a highly selective detection. These
techniques, however, can be developed and tested under less
stringent experimental conditions (with higher concentra-
tions) to investigate behaviors, fates and effects.
Table 3 Description of different types of size averages, with equations defining them and methods that are deriving such average sizes
Type of size average Applies to method Equation
Number average: size average of numbers of particles
within a certain size class
Microscopy, LIBD
d
n
¼
P
i
n
i
d
i
P
i
n
i
Mass or volume average: size average of volume of particles
within a certain size class
FFF and SEC with most detection
methods, CFF
d
v
¼
P
i
V
i
d
i
P
i
V
i
Z-average size, an intensity weighted average attributed to certain methods Dynamic light scattering
d
n
¼
P
i
n
i
d
6
i
P
i
n
i
d
5
i
Table 4 Specifications of methods for analysis and characterization of nanoparticles
Method Approximate size
range (nm)
Limit of detection
a
Single particle or particle
population methods
Level of sample
perturbation
AFM 0.5 to [1000 ppb–ppm sp Medium
BET 1 to [1000 Dry powder pp High
Centrifugation 10 to [1000 Detection dependant pp Low
Dialysis 0.5–100 Detection dependant pp Low
DLS 3 to [1000 ppm pp Minimum
Electrophoresis 3 to [1000 ppm pp Minimum
EM-EELS/-EDX Analysis spot size: *1 nm ppm in single particle sp High
ESEM 40 to [1000 ppb–ppm sp Medium
ES-MS \3 ppb pp Medium
FFF Flow FFF: 1–1000 Detection dependant; UV: ppm,
Fluo&ICP-MS: ppb
pp Low
Sed FFF: 50–1000
HDC 5–1200 Detection dependant pp Low
ICP-MS Depends on fractionation ppt–ppb pp
LIBD 5 to [1000 ppt sp Minimum
Microfiltration 100 to [1000 Detection dependant pp Low-medium
SEC 0.5–10 Detection dependant pp Medium
SEM 10 to [1000 ppb–ppm sp High
SLS 50 to [1000 pp Minimum
TEM/HR-TEM 1 to [1000 ppb–ppm sp High
TEM-SAED Analysis spot size: 1 nm sp High
Spectrometry ppb–ppm pp Minimum
Turbidimetry/nephelometry 50 to [1000 ppb–ppm pp Minimum
Ultrafiltration 1–30 Detection dependant pp Medium
WetSEM 50 to [1000 ppm sp Low
WetSTEM ppm sp Low
XRD 0.5 to [1000 Dry powder pp High
a
For comparison mass concentration limit of detection for 100 nm particles are estimated
Nanoparticle analysis and characterization methods 347
123
Dispersion, sampling and sample handling
Dispersion of nanoparticles for both exposure
and effect assessments
Colloidal systems are dynamic non-equilibrium systems and
are often sensitive to physical or chemical disturbances
(Filella 2007). Sampling and laboratory procedures (e.g.,
pumping, mixing, etc.) that introduce shear forces are likely
to perturb the dispersion state of ENPs, possibly leading to
either further aggregation, or to partial disruption of existing
aggregation. The presence of natural organic matter and
natural nanoparticles further complicates the situation. It is
important to be aware of and characterize the interaction of
the ENP with the natural material. It is equally important to
compensate for any background material of the same com-
position as the ENP. Background levels of identical
composition can be present for TiO
2
, SiO
2
but also for car-
bon-based nanoparticles. Geological studies, using primarily
transmission electron microscopy (TEM) to visualise the
materials, have reported fullerenes in geological formations
dating back 1.85 billion years (Becker et al. 1994), and
CNTs together with fullerene–like structures in a Greenland
ice core dated at approximately 10,000 years old (Murr et al.
2004). Given their reactivity, this is surprising (Taylor 2006),
but infers that these carbon-based nanoparticles have natural
as well as engineered origins.
In the case of ecotoxicological exposures to carbon
nanoparticles, the preparation and characterisation of aque-
ous fullerene suspensions is especially challenging owing to
their low solubilities. Fortner et al. (2005) describe nano-
aggregate formation of C60 fullerenes in water. Particle sizes
within the aggregates are, however, dependant on formation
parameters including pH, ionic strength and even the mixing
rates. The properties of the aggregates are different from the
pristine particles. Coupled with the fact that fullerenes oxi-
dise (Fig. 1), ecotoxicological exposure techniques are
rendered highly complex. For carbon nanotubes (CNTs),
their extremely low solubility in water, variable sizes of the
particles, small diameters and the complexity of aggregates
formed render dosing and particulate characterisations
extremely difficult in aqueous exposure experiments. Now-
ack and Bucheli (2007) describe a standard procedure for
solubilising CNT through cutting the tubes by sonication,
and hydroxylation of the ends and damaged regions using
strong acid. Other treatments to disperse the materials are
reported using surfactants (Jiang et al. 2003) and biopoly-
mers, including humic and fulvic acids (Hyung et al. 2007).
Table 5 Analytical chemistry, metrology and particle size analysis definitions
Term Definition
Metric The property that is being quantified
Analyte The compound or specie that is being quantified
Limit of detection The lowest concentration that can be distinguished from the background, typ defined as 3*Stdev (blank measurements)
Precision The statistical spread of values in a measurement series
Accuracy The exactness of the averaged measurements related to the true value
Measurement
uncertainty
The accumulated uncertainty incl. method, lab, between days and between lab biases
Method validation Experimental proof that the method conforms according to the specifications
Reference material A material or substance that is sufficiently homogeneous for its property values to be used for calibration of
instruments or assessment of methods
Certified reference
material
A reference material that is accompanied by a certificate that specifies the traceability of the CRM and associated
uncertainty
Control sample Within laboratory quality control over time and between interlaboratory comparisons
Interlaboratory
comparison
A blind test between participating laboratories to quantify deviation from true or reference value
Number based
concentration
Determinations of number of particles per unit volume or mass
Mass based
concentration
Determinations of mass of particles per unit volume or mass
Number average based
size
The size average of numbers of particles within a certain size class:
d
n
¼
P
i
n
i
d
i
P
i
n
i
Volume average based
size
The size average of volume of particles within a certain size class:
d
v
¼
P
i
V
i
d
i
P
i
V
i
Z-average based size A light scattering based average:
d
n
¼
P
i
n
i
d
3
i
P
i
n
i
d
2
i
Polydispersity index Weight average size/number average size
348 M. Hassello
¨
v et al.
123
Treatments to facilitate dispersion must, however, be
accounted for in interpretation of toxic response and how
environmental relevance may be affected.
Sampling
Due to the unstable nature of colloidal nanoparticle dis-
persions it is preferable to use in situ analyses, but these
methods are rarely available (Lead and Wilkinson 2006).
The second choice is to apply methodologies that cause
minimum perturbation from sampling to analysis. An
example of such techniques are the probing of dispersions
with electromagnetic radiation (e.g., light, X-rays or neu-
trons) where the scattering/absorption patterns can be
related to physical properties of the particles, as will be
described below.
Sample contamination and loss
Sampling of nanoparticles should generally be feasible
with most standard sampling protocols, but the handling
procedures differ from many other chemicals. Samples of
colloids from surface waters are often collected in bottles
that have been selected for minimum adsorption and con-
tamination, e.g., plastics, especially fluoroplastics, for
inorganic colloids or metal analysis and glass for analysis
of organic trace constituents (Hall 1998). Since engineered
nanoparticles may consist of e.g., an inorganic core with an
organic coating or surfactants, conventional material
selections may have to be revised. Further, the nanoparticle
surface charge and possible charges on the bottle walls of
both plastic and glass at the specific pH should be taken
into account. Consequently, for engineered nanoparticles,
adsorption to sample bottles needs to be investigated for
both inorganic and carbon-based nanoparticles on a case-
by-case basis until new experience-based knowledge has
been accrued. Similar concerns apply to all other materials
to which the sample is being exposed (e.g., tubing, filter
materials, pipettes, amongst others).
Extracting inorganic nanoparticles from soil and
sediment
Examining ENP in soils and sediments have the same
limitations as for water samples, with the additional com-
plication of much higher quantities of natural solids, many
of which are in the same size range as the ENP. Dispersion
methods for releasing natural nanomaterials from the solid
matrix, such as sonication and chemical dispersants
(hexametaphosphate, detergents, etc.) will likely release
the ENP to the solution phase, but the physicochemical
state of ENP will be likely to change (e.g., break-up of
flocs). These protocols are reported in the soil literature
(Gee and Bauder 1986). The separation of nanoparticles
from soil suspensions or sediment slurries are difficult, and
are prone to artifacts. As a general suggestion, centrifu-
gation is generally less perturbing than filtration (Gimbert
et al. 2005, 2006), but the differential settling during cen-
trifugation can also induce aggregation. This is further
discussed in the ‘Prefractionation’ section below. The
challenge then remains to discriminate between natural and
engineered nanoparticles.
Extracting carbon-based nanoparticles from water, soil
and sediment
Pristine fullerenes are comparatively soluble in organic
solvents such as toluene and can be extracted from media
(including water) into solvent (Fortner et al. 2005). In the
case of CNTs (both single and multi-walled), Nowack and
Bucheli (2007) summarise that no method currently exists
for their quantification in natural media. Indeed, CNTs
have low solubility, even in organic solvents.
Prefractionation
Environmental samples often contain complex mixtures of
particles of different size classes, composition, shapes and
are of biotic and/or abiotic origin. In order to study nano-
particles, it is often necessary to first reduce the complexity
using a course prefractionation. The prefractionation can be
based on settling, centrifugation or filtration. Settling or
centrifugation is only effective in removing particles that
have a settling velocity that dominates over their Brownian
motion. The settling velocity depends on the particle vol-
ume, shape, and their density difference with respect to
water. Therefore, settling or centrifugation is more efficient
in removing more dense mineral particles than it is for
algae and other organic particles. Centrifugation is a min-
imum perturbation prefractionation technique, but settling
particles can scavenge other smaller particles due to the
differential settling velocities.
Microfiltration, with pore sizes generally greater than
0.1 lm, is the most common prefractionation technique,
due to its simplicity of operation. However, common
‘dead-end’ filtration is prone to many artifacts, e.g.,
nanoparticle deposition, membrane concentration polari-
zation, and filter cake formation (Buffle et al. 1992;
Morrison and Benoit 2001).
Nanoparticles can be deposited on the membrane sur-
face due to collision or electrostatic attraction. Particles
smaller than the pore size can be transported through the
membrane more slowly than the liquid, due to electrostatic
repulsion within the pores. This causes concentration
polarization (build up of higher particle concentration in
the membranes diffusive boundary layer) which leads to
Nanoparticle analysis and characterization methods 349
123
higher collision rates between particles and consequently
aggregation. Aggregates or attached particles on the
membranes, provides more efficient trapping of nanopar-
ticles and their aggregates. This leads to formation of a
filter cake and the effective pore size decreases severely; in
other words the filter clogs.
These problems are especially severe for non-stabilized
nanoparticles, e.g., those that lack hydrophilic surfaces.
Therefore, filtration of engineered nanoparticle suspensions
should be critically evaluated in terms of the scavenging of
nanoparticles and, as a consequence, changing the size
distribution.
Fractionation by ultrafiltration, nanofiltration
and dialysis
Fractionation by membranes can either be done by apply-
ing a pressure to overcome the pressure drop across a
membrane that sieves molecules or particles according to
their size as in ultrafiltration or it can be done by letting
solutes equilibrate across the membrane as in dialysis. The
microfiltration artifacts mentioned above become greater as
the pore size of the filter decreases (ultrafiltration and
nanofiltration). This is especially critical where membranes
are used as macromolecular sieves. In order to reduce the
diffusive boundary layer over the membrane, and thereby
minimize the concentration polarization over the mem-
brane, cross-flow (or tangential) filtration (CFF) has been
developed. In CFF the sample is recirculated (or stirred) in
a reservoir on top of the membrane. A fraction of the
sample with components smaller than the pore size, will
pass through the membrane (to yield the permeate) in each
cycle. By measuring the concentration of analyte in both
the initial sample, the retentate (the fraction not passing
through the membrane) and the permeate, it is possible to
calculate the concentrations of analyte in the fractions
smaller and larger than the membrane pore size. The per-
formance of crossflow ultrafiltration has been extensively
evaluated for natural colloids and reveals that the mem-
brane type, membrane manufacturer, and operating
conditions, have large influences on the fractionation
results and recoveries obtained (Guo et al. 2000; Larsson
et al. 2002; Liu and Lead 2006). Therefore, crossflow
ultrafiltration should be appropriately tested and evaluated
prior to application to ENPs. Ultrafiltration is a preparative
size fractionation method that can be scaled to process
large sample volumes and produce large quantities of iso-
lated nanomaterials. Although it is limited to two fractions
(above and below the membrane pore size), multi stage
filtrations can allow for a crude size fractionation, however,
this is extremely labor and time intensive. When the
membrane pore size is below *1 nm, the method is
typically defined as nanofiltration. Nanofiltration is usually
applied to the separation of molecules from salts and could
potentially be applied to separate nanoparticles from their
dissolved counterparts.
Dialysis is an ultra- or nanofiltration method that oper-
ates on diffusion of solutes across a membrane that arises
from concentration gradients and osmotic pressure instead
of pressure driven filtration (as is the case in CFF). Dialysis
is a very mild fractionation method and it can be used to
separate truly dissolved components (ions and small mol-
ecules) from their nanoparticle counterparts. Dialysis has
been used to study nanoparticle-solute sorption behavior as
well as nanoparticle dissolution, where the aqueous coun-
terparts will diffuse across the dialysis membrane (Franklin
et al. 2007). However, dialysis usually utilizes deionized or
distilled water as an acceptor solution. This may promote
dissolution or ionic strength changes which will lead to
changes in dispersion state.
Field-flow fractionation, size exclusion
and hydrodynamic chromatography
Field-Flow Fractionation (FFF) is a mild chromatography-
like size-fractionating method that differs from chroma-
tography in that it does not utilize a stationary phase. The
most common FFF sub-technique is Flow FFF, which is
discussed here. Flow FFF separates nanoparticles accord-
ing to their particle size by virtue of their diffusion
coefficients in a very thin open channel (Giddings 1993;
Hassello
¨
v et al. 2007; Schimpf et al. 2000). The separation
principle relies on the combination of an applied field and
longitudinal carrier flow. The field acts perpendicular to the
length of the separation channel and causes the nanopar-
ticles to move towards the accumulation wall.
Nanoparticles form a cloud whose thickness is given by the
particles’ ability to oppose (generally through diffusion)
the force of the field. Smaller particles will not be affected
to the same extent as larger particles, and hence the smaller
particles elevate higher in the channel. Perpendicular to the
field, along the channel, the laminar separation flow is
acting on the nanoparticles. The parabolic shape of the
laminar flow velocity in the channel implies that particles
traveling nearer to the middle of the channel move faster
than particles traveling closer to the channel walls. Con-
sequently, the smaller particles, having higher extending
clouds, on average, travel faster than the larger particles,
resulting in fractionation of the sample that provides a
continuous size distribution. To monitor the size distribu-
tions, the FFF needs to be coupled to a detector that
responds to the nanoparticle number or mass concentration.
Examples include: UV absorbance, light scattering (von
der Kammer et al. 2005b), or elemental detectors such as
350 M. Hassello
¨
v et al.
123
ICP-MS (Hassello
¨
v et al. 1999; Ranville et al. 1999;
Jackson et al. 2005). The latter detector is very useful for
characterizing metal-containing nanoparticles, an example
being given in Fig. 2. Depending on the type of detector
used, different kinds of size dependant information of the
sample is achieved. One great advantage with FFF, com-
pared to other fractionation methods, is that the retention
time is directly proportional to nanoparticle physical
properties. Retention in FFF is expressed as the retention
ratio (R) given by
R ¼
t
0
t
r
ð1Þ
where t
0
is the void time and t
r
is the sample retention time.
For highly retained components, R can be approximated by
R 6k ð2Þ
while R can be estimated as follows for intermediate
retention
R ¼ 6k coth
1
2k

2k

ð3Þ
The fundamental retention parameter (k) is defined as
the mean distance of the component from the wall (l)
divided by the channel thickness (w).
k ¼
l
w
¼
D
Uw
ð4Þ
Channel thickness is calculated from experimentally
determined channel volumes, since the actual channel
thickness may differ from the manufacturer’s specifica-
tions. Estimates of k from experimental determinations of
R allow calculation of the diffusion coefficient (D). It is
important to note that the fundamental measurement made
by Flow FFF is the diffusion coefficient. In the techniques
of Flow FFF, diffusion coefficients can be used to deter-
mine hydrodynamic diameter. In Sedimentation FFF
buoyant mass or equivalent spherical diameter can be
determined (Giddings 1993).
The most critical factor in Flow FFF analysis is the
choice of membrane and the carrier composition optimi-
zation. The particles should travel through the fractionation
channel in close vicinity to the membrane without aggre-
gating, adsorbing to the membrane or having inter-particle
repulsion. This is generally accomplished for the complex
natural samples by controlling the electrostatic repulsion
and steric stabilization by a combination of suitable ionic
strength (typically 0–20 mM monovalent salt) and a sur-
factant (e.g., 0.05% sodium dodecyl sulphate) (Hassello
¨
v
et al. 2007). FFF has been successfully applied to a wide
range of synthetic nanoparticles (e.g., SiO
2
, TiO
2
, ZrO
2
,
Au, Ag, carbon black, pigments, Teflon, carbon nanotubes,
soot particles) (Schimpf et al. 2000).
Another size fractionation method is size exclusion
chromatography (SEC) where a particle or macromolecule
mixture is passed through a column with a porous packing
material with a distribution of pore sizes in the range of
particles to be fractionated (Barth and Boyes 1992). The
particles are separated according to their hydrodynamic
volume (size and shape) by their ability to enter the porous
structure of the packing materials. Particles that are larger
enter pores to a lesser extent than the smaller particles.
Each SEC column has a certain operating size (or molar
mass) window, and the first eluting larger particles (all at
once) are those outside the operating window, then come
the fractionated particles and then the ‘salt peak’ ions and
molecules that have passed through the complete pore
volume. Size exclusion chromatography has been applied
to both carbon nanotubes and fullerenes, as described in a
later section, to natural organic and inorganic nanomate-
rials (Perminova et al. 2003; Vogl and Heumann 1997;
Jackson et al. 2005).
Hydrodynamic chromatography (HDC) is another size
fractionation method that is carried out in narrow open
capillaries, or in wider capillaries with non-porous packing
materials that essentially form capillary routes. Due to the
size, the center of mass cannot approach the walls infinitely
and therefore a smaller particle can approach the wall to a
larger extent than can a large one. Therefore, the elution
order is the same as in SEC and also in the steric mode of
FFF. The separation efficiency of HDC is very poor, but
the operating size range is very good. HDC has been suc-
cessfully applied for the fractionation of nanoparticles
(Williams et al. 2002; Tiede unpublished results).
Chromatographic analyses of carbon nanoparticles
Many conventional techniques have been used to analyse
fullerene solutions including UV–vis spectrophotometry,
0
10
20
30
40
50
60
2400180012006000
Retention time (sec.)
metal (µg l
-1
)
0
Hydrodynamic diameter (nm)
Relative Fluoresence
82Se
111Cd
Fl
40 80 120 160
Fig. 2 Representative FFF fractogram of a CdSe quantum dot using
on-line fluorescence and ICP-MS detection
Nanoparticle analysis and characterization methods 351
123
infrared spectroscopy, nuclear magnetic resonance, and
mass spectrometry, frequently coupled to high performance
liquid chromatography (HPLC) (Andrievsky et al. 2002;
Fortner et al. 2005; Isaacson et al. 2007; Nowack and Buc-
heli 2007; Treubig and Brown 2002). For HPLC, octadecyl
silane (ODS) stationary phases are most commonly selected
with elution using solvents such as toluene or toluene:ace-
tonitrile mixtures (Treubig and Brown 2002). When UV–vis
absorbance detection is used, 325 nm is the wavelength
typically selected. Alternatively, gel permeation chroma-
tography can be used, for example using Agilent PL gel
10 lm 50 A with toluene elution (Readman and Frickers,
unpublished data). Size exclusion chromatography has also
been applied to characterise CNTs (Duesberg et al. 1998).
Light scattering techniques
Light scattering is a very commonly used method to
determine particle size (Schurtenberger and Newman
1993). The electromagnetic radiation of the incident pho-
tons induces an oscillating dipole in the particle electron
cloud. As the dipole changes, electromagnetic radiation is
scattered in all directions. The light source could be laser
light, X-rays or neutrons, each of which enables probing at
different size ranges and particle compositions. Discussion
will mainly be limited to describing methods utilizing laser
light, since these are the most readily available methods to
be used in particle characterization for ecotoxicology.
Dynamic light scattering
In dynamic light scattering (DLS), also called photon
correlation spectroscopy or quasielastic light scattering,
fluctuations in the scattered light that depend on particle
diffusion is utilized. The fluctuations originate from the
Brownian motion of the particles and from the fact that
neighboring particles can have constructive or destructive
interference of the scattered light intensity in a certain
direction. In the DLS instrument the intensity is measured
over very short time periods (dt) and then it is possible to
compare (correlate) the intensity at time t0 with time
t0 + dt (in the order of micro-milliseconds). Smaller par-
ticles (with faster diffusion) lose the correlation (the
memory of their previous position) more rapidly than lar-
ger particles. The scattering intensity is plotted as an
autocorrelation function:
g sðÞ¼G sðÞI
hi
2
=c
1=2
¼ Ae
2Cs
ð5Þ
where G(s) is the field autocorrelation function, I
hi
2
is the
base line and c is the coherence factor, expressing the
efficiency of the photon collection. A is an instrument-
specific constant, C is the decay rate and s the delay time.
C can be converted to the diffusion coefficient, D, using the
relation:
D ¼ C/q
2
ð6Þ
where q is the wave vector, which can be described by the
following relation:
q ¼ 4pg sin p=4ðÞ=k ð7Þ
where g is the refractive index of the solvent and k is the
wavelength of the incident light. If the diffusion coefficient
is known, the hydrodynamic radius, R
h
, can be calculated
from the Stokes–Einstein equation:
R
h
¼ kT= 6pgD ð8Þ
where k is Bolzmann’s constant and T is the absolute
temperature.
The advantages of DLS are: the rapid and simple
operation, readily available equipment, and minimum
perturbation of the sample (Ledin et al. 1994). The limi-
tations are the interpretation, especially for polydisperse
systems, and critical review of the data obtained (Filella
et al. 1997). DLS gives an intensity weighted correlation
function that can be converted to an intensity weighted
(z-average) diffusion coefficient.
For d \ k/20, then the scattering intensity, I * d
6
,
according to the Rayleigh approximation, while for k/
20 \ d [ *k then I * d
2
(Debye approximation). The
strong particle size dependence of the scattering intensity
will bias the measured size, as a small amount of large
particles will have such a large influence that smaller
particles will be neglected. Consider a sample with two
particle sizes, d: 3 and 30 nm, of equal particle number
concentrations. The volume concentration will be 1,000
times larger in the 30 nm particles due to the geometrical
formula of a sphere, but according to the Rayleigh
approximation, the scattering intensity will be 10
6
times
stronger for the 30 nm particle compared to the 3 nm
particle. For even larger particles, the response difference
will be enormous. Consequently, even the smallest fraction
of dust or other micrometer-sized particles will ruin the
signal from the nanoparticles.
For multimodal size distributions (multi component
mixtures), the conversion of the autocorrelation function to
diffusion coefficient is an ill-posed mathematical problem,
where small variations can give large deviations in the
output. For this reason, but more importantly due to the fact
that the signal from larger particles dominates over smaller
ones, a general rule is that DLS is not suitable for samples
with polydispersity index above *1.5–1.7.
Since DLS measures diffusion coefficients, and that all
size calculations are based on assumptions that the Stokes–
Einstein relation (Eq. 8) holds, it is essential to validate that
352 M. Hassello
¨
v et al.
123
the diffusion coefficient measured is the undisturbed self-
diffusion coefficient. For charged nanoparticles, electro-
static forces between particles have an effect on the
diffusive behavior. This effect is concentration dependant,
and the upper boundary occurs when the nanoparticle gets
entrapped by forces from their close neighbors, the point of
so-called gel-formation. By dilution of the sample to the
greatest possible extent, while remaining above the detec-
tion limit, and extrapolation of the measured diffusion
coefficient to infinite dilution, the unperturbed diffusion
coefficient can be estimated. This value is one that can
most reliably be used to calculate size in the Stoke Einstein
equation. However, dilution of a sample will change its
diffusion behavior and aggregation state. If primary parti-
cle size is not the goal, but rather to characterize the
dispersion state in a sample, then it is more relevant to not
dilute the sample, reporting diffusion coefficients only,
rather than size.
It should also be noted that the derived data from DLS
are intensity based distributions or averages, and mathe-
matical conversions to volume or number distributions
should only be provided with good knowledge of the par-
ticle shapes, polydispersity and underlying assumptions
(Finsy 1994). Although dynamic light scattering does not
provide full characterization of nanoparticle dispersion, it
is very valuable to, for example, monitor aggregation
behavior.
Static light scattering
Static light scattering (SLS), also called multi angle (laser)
light scattering (MALS or MALLS), provides measure-
ment of physical properties that are derived from the
angular dependency of light scattered by a particle. This is
due to the fact that a particle of a certain size generates
destructive and constructive interferences at certain angles.
Time averaged scattering intensities are measured at sev-
eral angles to derive any number of several size parameters
including the particle size, root mean square radius of
gyration (R
g
.), which is the root mean square distance of
point masses in a particle from its center of gravity. Con-
sequently SLS relates to the particle structure and
morphology and can therefore be used in combination with
DLS to give information of particle shape factors. There
are several important assumptions in SLS theory for dif-
ferent analytical solutions. The most used is called
Rayleigh–Gans–Debye approximation (Schurtenberger and
Newman 1993). For these approximations, the refractive
index difference between the particle and solvent should be
negligible, the concentration of particles approaches zero,
and no light absorption by the particles occurs.
Both dynamic and static light scattering polydisperse
samples impose limitations on these methods. Therefore,
it has been shown to be beneficial to couple light scat-
tering detectors online to a fractionation method such as
FFF or SEC (von der Kammer et al. 2005b; Wyatt
1998). With this combination, independent size distribu-
tions can be derived from the two methods and thereby,
from comparison of the two results, distributions of
particle shape factors can be estimated (von der Kammer
2005).
Nephelometry
Turbidity, or nephelometry, is a particle concentration
measurement that utilizes scattering of light at 90° or
sometimes 180°, with respect to the light source. The light
source can be a laser or monochromatic light. The equip-
ment is very simple and can be portable or even in situ, but
the relationship between the concentration and particle
concentration is not trivial. The light scattering intensity is,
as mentioned above, strongly dependent on particle size,
and also on other parameters such as the refractive index
difference between the particles and the suspension media.
Therefore, in quantitative analysis, turbidity measurements
should only be used for well-defined particles of fairly
narrow size distributions and complemented by calibration
with other techniques (e.g., gravimetry). For dispersed
nanoparticles, turbidity is fairly insensitive, and is less
suitable than for monitoring aggregation.
Nephelometry has also been used as a chromatographic
particle concentration detector (von der Kammer et al.
2005a).
Laser induced breakdown detection
Laser induced breakdown detection (LIBD) is based on the
fact that when a solid nanoparticle passes through the focal
volume of a focused, pulsed laser, the power density
required to induce breakdown of the dielectric properties of
the water is lower than for pure water (Kim and Walther
2007). If the laser energy is correctly tuned, plasma for-
mation will only occur when a nanoparticle passes through
the focal volume of the optical cell. The plasma formation,
or breakdown, is detected with either a piezo-electric
crystal attached to the cuvette, or with a CCD camera
synchronized with the laser pulse. The parameter measured
is the breakdown probability (BP). Since BP for a given
laser energy depends both on particle concentration and on
size, it is necessary to elucidate both. The most common
mode is to tune the laser pulse energy and measure the BP
of the sample, and do the same for a set of calibration
standards of known size at different concentrations. The BP
for larger nanoparticles has a threshold (increased from
zero probability) at lower laser energies than smaller
nanoparticles. The BP-laser energy curves have different
Nanoparticle analysis and characterization methods 353
123
slopes depending on the concentrations that are also given
from the calibration standards.
The main advantage of LIBD is that it is extremely
sensitive even to small nanoparticles with detection limits
in the ppt (ng dm
-3
) range. In fact, LIBD is so sensitive
that most samples have to be diluted in order to not saturate
the breakdown probabilities.
The main disadvantages are that LIBD cannot discrim-
inate between different types of nanoparticles and even
more seriously, that different nanoparticle compositions
have different breakdown probabilities (instrument
responses). Therefore it is not possible to use one set of
calibration standards for different types of nanoparticles.
LIBD is a specialized technique that is not yet commer-
cially available.
Spectroscopic analysis and characterization
Certain classes of nanoparticles demonstrate strong fluo-
rescence and this property is utilized in many fields such as
medical imaging, immunoassays, photonics, amongst oth-
ers (Bailey et al. 2004). Quantum dots (QDs) are composed
of semi-conductor materials, for example CdSe, CdS,
CdTe, and are highly fluorescent. These particles can be
characterized by either their absorption or fluorescence
emission spectra. The absorption spectra is broad over low
wavelengths but displays a sharp peak, called the first
exciton peak at the upper wavelength of the absorption
spectra. This peak is generally in the order of 20–50 nm
lower in wavelength than the emission peak. The position
of this absorption peak can be correlated to the particle size
and is commonly used to monitor size in QD synthesis (Yu
et al. 2003). The emission peak tends to be fairly narrow,
on the order of 50 nm, with the wavelength being highly
sensitive to nanoparticle size. Measurement of fluorescence
spectra can thus also be used to determine particle size. In
natural systems, natural fluorophores contained in humic
substances and biological cells may interfere with these
determinations. Non-fluorescent nanoparticles such as sil-
ica can be labeled with dyes to impart fluorescence. In
some cases the fluorescence of the dye can be enhanced by
the presence of a second dye that can contribute its exciton
energy through a radiation-less transfer.
Quantitation of particle concentrations can be performed
using absorption or fluorescence if the optical constants of
the particles are known. For example, extinction coeffi-
cients for the first exciton peak of some QDs were
determined by Yu et al. (2003). It is yet to be determined
how significantly background absorption from natural
occurring materials in water will limit the usefulness of
UV–vis absorption for nanoparticle quantitation in aquatic
systems.
Both UV–vis absorption and fluorescence can be used as
online detectors for chromatography and FFF systems. The
extremely bright fluorescence of some nanoparticles should
provide low detection limits for these techniques. Figure 2
shows an example of the use of online fluorescence
detection with FFF for a CdSe quantum dot.
Fluorescence microscopy gives spatial information and
has been very useful in looking at the distribution of
nanoparticles in cells and organisms. For example, uptake
of QDs into the guts of filter feeding organisms is clearly
observable using fluorescence microscopy.
For naturally fluorescent material or labeled macro-
molecules, fluorescence correlation spectroscopy within
the focal point of a laser confocal microscope, have been
successfully applied to determine the diffusion coefficients
of these materials (Lead et al. 2000b). The principle is
similar to dynamic light scattering (also called photon
correlation spectroscopy), but the sensitivity is much better
for small (fluorescent) particles. The method should be
very suitable for studies of QDs in environmental media.
In describing the UV–vis absorption spectra of metal
NPs, the term surface plasmon is used, which describes the
oscillating electron clouds present at the metal-solution
interface. Particle size strongly affects the absorption
spectra through quantum confinement effects that are
important at the nanometer scale of materials. The smaller
the particle size, the lower the wavelength of light absor-
bed. Aggregation of NP results in band broadening and red
shifting of surface plasmon band and has been used to
study the effect of electrolytes on metal NP stability (Aryal
et al. 2006).
Particle shape characterization of metal NP is also
possible from examination of surface plasmons. While
spherical gold and silver NP have strong surface plasmon
bands at about 520 and 400 nm respectively, nanorods of
these metals show two bands, a red-shifted long-axis band
and a blue shifted short-axis band. The wavelength of the
long axis band is particularly sensitive to particle aspect
ratio. It has also been noted that Au nanorods have 106
stronger fluorescence than spherical Au NPs (Link and El-
Sayed 1999). Consequently, surface plasmon effects can be
used to study particle-particle interactions since the aspect
ratio changes when to single particles come close together.
Electron microscopy and atomic force microscopy
There are several powerful microscopy techniques that can
provide images of nanoparticle systems as well as addi-
tional information on elemental composition, structure and
even charges or force measurements. Microscopy methods
are all single particle methods, that is the data does not
arise from an ensemble of particles such as is the case with
354 M. Hassello
¨
v et al.
123
light scattering. This enables information to be collected on
each particle free from interferences from other particles or
background solutes. This gives good information on par-
ticle processes that sometimes cannot be obtained with
bulk analysis (Mavrocordatos et al. 2007). However, it also
means that even though a quantitative measurement with
sometimes fairly good accuracy can be achieved on a
single particle, it is only by counting and measuring
enough particles (of a certain type or in a certain size
range) that good enough counting statistics of the complete
sample can be obtained. This is needed in order to deliver a
quantitative analysis or characterization of the sample.
Sizing with microscopy means that an average size mea-
sured on a certain number of particles are a number
average, and in order to measure an accurate size distri-
bution of nanoparticles it is necessary to count and measure
thousands of particles in order to obtain a reliable counting
statistics of the very few larger nanoparticles in the size
distribution. The large particles in the distributions (or
aggregates), even if very few, can contribute substantially
to the volume or mass based distributions. In nanotech-
nology or material science this is not a problem since the
particles to be measured are of the same type and are of
similar size but when dispersed in water and mixed with
natural organic matter and natural nanoparticles it is
another story. Therefore we see a big need for automation
in electron microscopy and development of ‘smart’ image
analysis software that enables characterization of the mil-
lions of particles needed in each sample (Mavrocordatos
et al. 2007). With this said microscopy methods are very
powerful for imaging and process understanding but it
should be complemented with a particle population method
that is giving quantitative information on the sample.
Another common feature for all microscopy techniques
is that they require different levels of sample preparation. It
ranges from the mildest being drying of the particles to a
moist condition (AFM and ESEM) to a high-vacuum in
SEM and TEM. In some methods coating or staining the
sample is used. The transfer of the sample from its dis-
persed hydrated state to a dried high vacuum state often
means that the particle size distribution changes dramati-
cally. For example by evaporating a sample drop into
dryness (a common method) the particle concentration and
solute concentration increases drastically in the decreasing
volume of the drop before it finally evaporates. This leads
to aggregation of particles and precipitation of salts. Some
methods are used to preserve the hydrated state of particles
either by cryofixation, which is a rapid freezing so that the
water forms non-crystalline ice. Another method is
embedding the particles in some water-soluble resin that
fixes the water when it cures.
The three most common sample preparation methods for
natural colloids is drop deposition, adsorption deposition or
ultracentrifugation harvesting, and the methods have been
compared for AFM and electron microscopy respectively
(Balnois and Wilkinson 2002; Mavrocordatos et al. 2007).
Scanning electron microscopy
In the family of electron microscopy techniques the sample
is exposed to a high energy focused beam of electrons. In
scanning electron microscopy (SEM) the interaction of the
beam with the particle surface are scanned over the sample
and measured as secondary electrons (most common), or
backscattered electrons or X-ray photons. Due to the high
depth of field in SEM a three dimensional appearance can
be obtained. The sample needs to be conductively coated
with gold or graphite and maintained under ultrahigh
vacuum in order not to have the secondary electrons
interact with gas molecules. The substrate is typically a
filter membrane or a conducting grid.
Environmental scanning electron microscopy and
related techniques
Due to the problems with morphological changes of the
particles associated with the transfer to high vacuum state,
environmental scanning electron microscopy (ESEM) was
developed, where the sample cell is separated from the
detector cell. This allows the sample to be measured under
variable pressure and humidity (in theory up to 100%) with
residual hydration water still on the particles. This water
layer also serves as a conductor on the surface so the
sample does not need to be conductively coated. The res-
olution is decreased (from *10 to *100 nm) due to the
interactions of the secondary electrons with the water vapor
molecules but there are less sample artifacts for example
from natural colloids (Doucet et al. 2005). ESEM still
allows analysis of the emitted X-rays. Wet STEM is a
method for scanning TEM analysis of a wet sample on a
TEM grid in an ESEM microscope utilizing dark-field
imaging conditions with a resolution of a few tenths of nm
(Bogner et al. 2005). A new sample capsule (WetSEM
TM
)
with electron transparent membranes provides an alterna-
tive to ESEM in ordinary SEM microscopes. The WetSEM
capsules allow imaging under liquid or moist conditions
(Thiberge et al. 2004). However, the loss of resolution is
considerable (partly due to diffusion of the particles), and
the membrane is sensitive to radiation damage, and only
particles close to the membranes are in focus.
Transmission electron microscopy
In transmission electron microscopy (TEM) the electron
beam is transmitted through a very thin specimen on a
conducting grid (e.g., copper grid with a thin resin,
Nanoparticle analysis and characterization methods 355
123
e.g., formvar). After the beam has been transmitted through
the sample and has interacted with the particles the non-
absorbed electrons are focused onto an imaging detector
(fluorescence screen or CCD camera). In TEM the particles
are shined through by the electron beam and the absor-
bance (image contrast) is both a function of the electron
density of the elements in a particle and the thickness of the
particles. Organic matter with only light elements needs to
be stained by a heavy metal cocktail in order to be visible.
High-resolution TEM is a method that can give subna-
nometer resolution and is used in material science to study
atom-by-atom structure. HR-TEM is a very demanding and
time-consuming method but it has been applied to detect
nanoparticle formation by bacteria or in geochemical pro-
cesses (Banfield and Navrotsky 2001; Suzuki et al. 2002).
TEM has also been applied to characterize carbon
nanoparticle dispersions in ecotoxicological exposure
experiments (Smith et al. 2007).
Electron microscopy microanalysis
For all electron microscopy methods mentioned here
analysis of spectral patterns of emitted X-rays (K, L & M
lines) for elemental composition of the particles can be
utilized if the microscopes are fitted with an energy dis-
persive X-ray spectrometer (EDX or sometimes EDS). The
spatial resolution can be even less than10 nm. The sensi-
tivity is best for heavier elements, so in reality it works best
for major elements of the particles and associated heavy
metals in fairly high concentrations. The measurement
uncertainty of EDX is generally *20% (Mavrocordatos
et al. 2004, 2007).
Electron energy loss spectrometry (EELS) is another
elemental composition method that can be applied in either
spectrometric mode or in imaging mode in TEM. In EELS
the loss of energies due to inelastic scattering processes
(e.g., inner shell ionizations) can be interpreted to which
elements that were causing the scattering. The energies lost
are specific for each element. The EELS results are more
difficult to interpret than EDX and works best for the
lighter elements (from carbon and up to zinc). EELS can
also be used to obtain additional chemical information
(e.g., redox states of transition metals).
Atomic force microscopy
Atomic Force Microscopy (AFM) is a subnanometer res-
olution method in the family of scanning probe
microscopy. It utilizes a cantilever with a very thin tip (tens
of nm), that is oscillating over the surface of the sample.
The oscillating movement (Z-axis) and the scanning over
the surface (X and Y-axis) is controlled by piezoelectric
actuators.
A laser-based balance can measure both repulsive (Pauli
principle) and attractive (van der Waals) forces between
the tip and the sample in the range 10
-7
to 10
-12
N. The
occurrence of these forces at different stages of the canti-
lever oscillation can be used to derive a separating distance
between the tip and the particles. The resulting images are
an atomic force topography. The substrates that the particle
samples are prepared on should be atomically flat (mica,
graphite or silicon wafers are examples of suitable sub-
strates). The preparation methods are typically drop
deposition, adsorption deposition or ultracentrifugation as
for electron microscopy, but in addition it is possible to
analyze samples under moist conditions or even in liquids,
which affords minimum perturbation. However, under
liquid conditions the particles are sometimes attracted to
the substrate (very weakly) and are moved around dis-
turbing the images. Another feature in AFM is that the
geometry of the tip compared to particle size gives that the
tip is starting to ‘feel’ the particle significantly before its
center has approached the particle periphery, and analo-
gously when the scanning tip is leaving the particle it feels
the particle forces too long. Therefore the lateral dimen-
sions are greatly overestimated, while the height
measurements are very accurate. This should be kept in
mind when interpreting AFM images, which means, e.g., a
carbon nanotube can give a height of 1 nm but a width of
up to 50 nm even though these should be the same. The
geometry of the tip should be decreased if small particles
are to be more accurately probed. The cantilever tip can be
set to contact the particles but lateral forces lead to
movement of particles, so a tapping mode or a non-contact
mode has been developed to just feel the forces above the
particles (Balnois et al. 2007). The latter has shown to be
more accurate for soft, compressible particles such as
humic acids. AFM is one of the most common nanome-
trology methods and has numerous applications (e.g., Lead
et al. 2005; Viguie et al. 2007).
In Fig. 3, a dispersed ZnO nanopowder sample (with
manufacturer stated size 50–70 nm) has been prepared
with adsorption deposition and analyzed with AFM, TEM
and SEM. The difference in visualization and size mea-
surements is clear. AFM and TEM show sintered
aggregates with primary particles in the size range pro-
vided by manufacturers, whilst SEM shows mainly larger
flakes of material with some nanoparticles on top. It is
likely that the sample preparation and vacuum-induced
changes can explain these differences.
Surface charge measurements
Colloidal nanoparticles develop surface charges in aqueous
solutions. The net surface charge, or surface potential, is
356 M. Hassello
¨
v et al.
123
one of the most important nanoparticle characteristics since
it describes to what extent the nanoparticle dispersion is
electrostatically stabilized by interparticle repulsion. Con-
sequently, ENP surface potential will have major influence
on their fate and behavior (Guzman et al. 2006; Hunter and
Liss 1979). However, it is not easy to directly measure the
surface potential but there is a simple method that measures
the so-called zeta potential, which is the potential at a
hydrodynamic slipping plane in the electrostatic double
layer of the particles as measured by electrophoresis. The
measured electrophoretic mobility can be converted to z
potential through Smoluchowski’s theories. The point of
zero charge (PZC) is the pH where negative and positive
charges are balanced, so there is no net charge on the
nanoparticles. At PZC there is generally maximum aggre-
gation taking place since the particles are allowed to come
in close contact so that attractive van der Waals forces can
act.
Surface area measurement
The Brunauer, Emmett, Teller (BET) (Brunauer et al.
1938) method is used to measure the specific surface area
of solids, which involves drying of a powder in vacuum
and then measuring (using a microbalance) the adsorption
of dinitrogen gas (assumed as a monolayer) on the surface
and in micropores. The BET method builds on the
assumption that N
2
has access to the complete surface of
the particles. Other variants of this method based on
adsorption of organic molecules (e.g., ethylene glycol
monoethyl ether, EGME) can be used (Hassello
¨
v et al.
2001). Dinitrogen gas gives higher surface areas than
EGME, probably due to greater access to smaller pores.
Crystal structure
X-ray diffraction (XRD) is a method of measuring inter-
particle spacings resulting from interference between
waves reflecting from different crystal planes. It is used in
mineralogy to determine crystal structure of mineral par-
ticles. For example XRD can be used to distinguish
between the anatase and rutile and amorphous phases of
TiO
2
nanoparticles. A dry sample needs to be prepared as a
thin film. Elemental composition of major elements can
also be obtained although the sensitivity is low compared
to other elemental analysis methods (e.g., ICPMS or AES).
Fig. 3 ZnO nanoparticle
powder (50–70 nm, Sigma
Aldrich UK), dispersed in
distilled water (*5mgl
-1
),
allowed to dry on silica and
imaged by AFM (1a and b),
TEM (2) and SEM (3) under
standard conditions
Nanoparticle analysis and characterization methods 357
123
It is also possible in TEM to measure the diffraction
patterns of single particles using a method called ‘‘Selected
area electron diffraction’ (SAD, or SAED). In SAD the user
can select an area of the sample with a small aperture and
only the electron diffraction pattern from that area will be
measured. This has benefits over XRD for heterogeneous
samples because it allows single particle characterization.
Difference in analysis of particulate
and nanoparticulate assemblages compared
to conventional analysis of solutes
For analysis of nanoparticle assemblages by bulk analytical
methods (in contrast to single particle analysis methods, e.g.,
microscopy) in whole samples or on fractions after sample
treatment (e.g., filtration or Field-Flow Fractionation), it is
necessary to recognize that for certain methods there may be
differences compared to more common analysis of dissolved
solutes (e.g., ions or molecules). In bulk analysis of a
nanoparticle dispersion the analytes mass concentration are
not homogeneously distributed, but rather as uniformly
distributed point masses. This is not a problem providing the
probed sample volume of the method is not approaching that
of single nanoparticles. But when analyzing samples with
environmentally relevant concentrations, with methods that
are probing a very small samples volume (e.g., a very rapid
measurement in a capillary or a fast flowing sample stream
such as in mass spectrometers) the measurement may
approach or enter a domain of single nanoparticle events.
The consequence is a noisier signal and if there is statistically
less than one particle per measurement then the recovery of
the determination decreases, which gives an erroneous
determination. Since the particle numbers (for the same
mass) decreases rapidly for larger particles, this issue is more
severe for them than for smaller particles. This is a well-
known phenomena in e.g., ICPMS analysis of micrometer
sized particles, and is called slurry nebulization. It needs to
be considered when the number concentration is low. Other
problems may be non-quantitative measurement of the par-
ticles, for example through incomplete atomization in
elemental analyses or non-transparent or shading effects in
spectroscopy. For slurry nebulization in ICP-AES or ICP-
MS, it has been found that the particle size is the dominating
factor to obtain complete atomization, where particles below
3–5 lm have been found to yield quantitative recoveries
compared to solutions (Ebdon et al. 1997; Santos and Nob-
rega 2006). The main reason for decreasing recoveries was
poor transport efficiencies in the nebulizer-spraychamber
system. This implies that for nanoparticles, incomplete
atomization should not be a problem but maybe for aggre-
gates particularly refractory materials such as carbides and
some oxides may also present problems.
Validation, measurement uncertainty and good
laboratory practices
In metrology and analytical chemistry, it is fundamental to
be able to report on the traceability of the acquired results.
Calibration standards used for quantification are generally
traceable to a primary national or international standard.
However, for nanoparticles the validity of these standards
has a shorter lifetime than most other standards and is more
sensitive to operating conditions. Nanoparticle standards,
or reference materials, exist both as suspensions and as
powders. Nanoparticle standards in suspension are gener-
ally labeled with expiry dates and instructions for storage.
Sometimes there are also instructions on how to further
dilute the standard in order to maintain its integrity. The
use of powdered nanoparticle standards does not include a
standardized procedure for dispersion of the nanoparticles.
To make the dispersion in each individual laboratory
increases the uncertainty of the original metric stated by
the manufacturer. Indeed, many metrics (e.g., size distri-
bution) are strongly dependant on how the dispersion was
made and in which media (pH, ionic strength and com-
position and presence of organic matter).
In addition to the nanometrology specific issues of method
validation relates to the normal quality control (QC) of any
analytical method (Table 5). The most important steps in
analytical QC are method validation and quantification of
measurement uncertainty. The method validation is simply
an experimental procedure to determine that the method and
procedures (standard or in-house developed) are complying
with the documented specifications (e.g., limit of detection,
linearity, determination of precision and accuracy and
robustness). One way of determining the accuracy is to use a
certified reference material (CRM) of the same type as the
samples and with documented property values within the
range of the method (Table 5). CRMs or NIST traceable size
standards are however very rare for nanoparticles as yet.
There exist reference materials with certified sizes for gold
and polystyrene colloids in the nanometer size range. More
reference materials are under development through inter-
national efforts. Testing the homogeneity, shelf life of a
reference material and carrying out all the analysis in order to
certify the material is very elaborate and expensive. In the
absence of CRMs there is also the possibility to use non-
certified materials to (test materials) to benchmark analytical
procedures and toxicity testing (Aitken et al. 2007).
Another option is to participate in interlaboratory compar-
isons were a blind sample is sent to many laboratories for
analysis, thereby affording a good indication of accuracy and
precision in the results. Interlaboratory comparisons are not yet
as common in nanometrology as in conventional analytical
chemistry where rigorous quality assurance protocols are fol-
lowed in order to achieve and maintain certified accreditation.
358 M. Hassello
¨
v et al.
123
There are, however, a few examples of informal interlabora-
tory comparisons on natural nanoparticles (Lead et al. 2000a)
and on engineered nanomaterials (Breil et al. 2002)which
have proved highly informative to the participants and for other
users of the same methodologies. A good daily routine is to
analyze a QC sample and plot that value into a control diagram
to monitor measurement uncertainty between interlaboratory
comparisons. The QC sample should be a sample that is stable
over time and that is as similar to the usual samples as possible.
Thus can method or instrument related problems in the labo-
ratory can be easily and quickly discovered.
Good laboratory practices in characterization of exposure/
effect experiments should include minimal sample perturba-
tion and determination of the dispersion-agglomeration state.
Dynamic light scattering fulfills these criteria, is a simple
measurement to perform, and is available in most academic
institutions. It is also a simple measurement to perform.
However, for the reasons described previously, the results
from DLS should not be over interpreted. DLS is primarily not
a size determination method as it measures scattering intensity
weighted diffusion coefficients. Thus, it is well suited to fol-
low initial stages of aggregation, but not to provide
nanoparticle sizes. For toxicity tests of nanoparticles, we
suggest to conduct a separate dispersion experiment under
optimum conditions as a reference to the dispersion behavior
in the effect media and during the course of the effect exper-
iment. This reference experiment with maximum dispersion
may include surfactants, co-solvents, certain ionic strength
and sonication. By comparing the results in the realistic effect/
exposure experiments with this reference experiment one can
obtain information on the degree of aggregation.
If competence and equipment is available a less biased
(but with slightly more perturbation) determination of the
size distribution can be achieved using e.g., Field-Flow
Fractionation. Microscopy (e.g., AFM, SEM or TEM) is
very powerful in imaging nanoparticles and aggregates, but
the aggregation state of the sample may have changed
during sample preparation.
Acknowledgements Hassello
¨
v thanks the Swedish Environmental
Research Council FORMAS and University of Gothenburg Nano-
particle platform for financial support. J. Readman acknowledges
partial support of his contribution through the UK Natural Environ-
ment Research Council Environmental Nanoscience Initiative (Grant
Reference Number: NE/E014321/1). Ranville acknowledges partial
support through EPA STAR Grant RD-83332401-0
References
Aitken RJ, Chaudhry MQ, Boxall ABA, Hull M (2006) Manufacture
and use of nanomaterials: current status in the UK and global
trends. Occup Med 56:300–306
Aitken RJ, Hankin SM, Tran CL, Donaldson K, Stone V, Cumpson P,
Johnstone J, Chaudhry Q, Cash S (2007) REFNANO: reference
materials for engineered nanoparticle toxicology and metrology.
DEFRA, UK. http://www.iom-world.org/pubs/REFNANOReport.
pdf
Andrievsky GV, Klochkov VK, Bordyuh AB, Dovbeshko GI (2002)
Comparative analysis of two aqueous-colloidal solutions of C-60
fullerene with help of FTIR reflectance and UV–vis spectros-
copy. Chem Phys Lett 364:8–17
Aryal S, Bahadur KCR, Bhattard N, Kim CK, Kim HY (2006) Study
of electrolyte induced aggregation of gold nanoparticles capped
by amino acids. J Colloid Interface Sci 299:191–197
Bailey RE, Smith AM, Nie S (2004) Quantum dots in biology and
medicine. Physica E 25:1–12
Balnois E, Wilkinson KJ (2002) Sample preparation techniques for the
observation of environmental biopolymers by atomic force micros-
copy. Colloids Surf Physicochem Eng Aspects 207:229–242
Balnois E, Papastavrou G, Wilkinson KJ (2007) Force microscopy
and force measurements of environmental colloids. In: Wilkin-
son KJ, Lead JR (eds) Environmental colloids and particles:
behaviour, structure and characterization. IUPAC series on
analytical and physical chemistry of environmental systems.
John Wiley and Sons, Chichester, pp 405–468
Banfield JF, Navrotsky A (eds) (2001) Nanoparticles and the
environment. Reviews in Mineralogy & Geochemistry, vol 44.
The Mineralogy Society of America, Washington, DC, p 349
Barth HG, Boyes BE (1992) Size exclusion chromatography. Anal
Chem 64:428R–442R
Barth HG, Flippen RB (1995) Particle size analysis. Anal Chem
67:257R–272R
Becker L, Bada JL, Winans RE, Hunt JE, Bunch TE, French BM
(1994) Fullerenes in the 1.85-billion-year-old Sudbury impact
structure. Science 265:642–645
Bogner A, Thollet G, Basset D, Jouneau PH, Gauthier C (2005) Wet
STEM: a new development in environmental SEM for imaging
nano-objects included in a liquid phase. Ultramicroscopy
104:290–301
Breil R, Fries T, Garnaes J, Haycocks J, Huser D, Joergensen J,
Kautek W, Koenders L, Kofod N, Koops KR, Korntner R,
Lindner B, Mirande W, Neubauer A, Peltonen J, Picotto GB,
Pisani M, Rothe H, Sahre M, Stedman M, Wilkening G (2002)
Intercomparison of scanning probe microscopes. Precis Eng-J Int
Soc Precis Eng Nanotechnol 26:296–305
Brunauer S, Emmet PH, Teller E (1938) Adsoprtion of gases in
multimolecular layers. J Am Chem Soc 60:309–319
Buffle J, Perret J, Newman J (1992) The use of filtration and
ultrafiltration for size fractionation of aquatic particles, colloids
and macromolecules. In: Buffle J, van Leeuwen HP (eds)
Environmental particles I. Lewis, Chelsea, pp 171–230
Colvin VL (2003) The potential environmental impact of engineered
nanomaterials. Nat Biotechnol 21:1166–1170
Crane M, Handy RD (2007) An assessment of regulatory testing
strategies and methods for characterizing the ecotoxicological
hazards of nanomaterials, Report for Defra, London, UK. Available
at: http://randd.defra.gov.uk/Document.aspx?DocumentID=2270
Doucet FJ, Lead JR, Maguire L, Achterberg EP, Millward GE (2005)
Visualisation of natural aquatic colloids and particles—a com-
parison of conventional high vacuum and environmental
scanning electron microscopy. J Environ Monit 7:115–121
Duesberg GS, Burghard M, Muster J, Philipp G, Roth S (1998)
Separation of carbon nanotubes by size exclusion chromatogra-
phy. Chem Commun 3:435–436
Ebdon L, Foulkes M, Sutton K (1997) Slurry nebulization in plasmas.
J Anal Atom Spectrom 12:213–229
EPA (2007) Nanotechnology white paper, U.S. Environmental Protection
Agency, Washington, DC. http://es.epa.gov/ncer/nano/publications/
whitepaper12022005.pdf
Federici G, Shaw BJ, Handy RD (2007) Toxicity of titanium dioxide
nanoparticles to rainbow trout (Oncorhynchus mykiss): gill
Nanoparticle analysis and characterization methods 359
123
injury, oxidative stress, and other physiological effects. Aquat
Toxicol 84:415–430
Filella M (2007) Colloidal properties of submicron particles in natural
waters. In: Wilkinson KJ, Lead JR (eds) Environmental colloids
and particles: behaviour, structure and characterization. IUPAC
series on analytical and physical chemistry of environmental
systems. John Wiley and Sons, Chichester, pp 17–93
Filella M, Zhang J, Newman ME, Buffle J (1997) Analytical
applications of photon correlation spectroscopy for size distri-
bution measurements of natural colloidal suspensions. Colloids
Surf A 120:27–46
Finsy R (1994) Particle sizing by quasi-elastic light-scattering. Adv
Colloid Interface Sci 52:79–143
Fortner JD, Lyon DY, Sayes CM, Boyd AM, Falkner JC, Hotze EM,
Alemany LB, Tao YJ, Guo W, Ausman KD, Colvin VL, Hughes
JB (2005) C60 in water: nanocrystal formation and microbial
response. Environ Sci Technol 39:4307–4316
Franklin NM, Rogers NJ, Apte SC, Batley GE, Gadd GE, Casey PS
(2007) Comparative toxicity of nanoparticulate ZnO, bulk ZnO,
and ZnCl
2
to a freshwater microalga (Pseudokirchneriella
subcapitata): the importance of particle solubility. Environ Sci
Technol 41:8484–8490
Gee GW, Bauder JW (1986) Particle size analysis, methods of soil
analysis. Part 1. Physical and mineralogical methods. Agronomy
monograph no. 9, 2nd edn. SSSA, Madison, Wisconsin, pp 383–411
Giddings JC (1993) Field-flow fractionation: analysis of macromo-
lecular, colloidal, and particulate matter. Science 260:1456–1465
Gimbert LJ, Haygarth PM, Beckett R, Worsfold PJ (2005) Compar-
ison of centrifugation and filtration techniques for the size
fractionation of colloidal material in soil suspensions using
sedimentation field-flow fractionation. Environ Sci Technol
39:1731–1735
Gimbert LJ, Haygarth PM, Beckett R, Worsfold PJ (2006) The
influence of sample preparation on observed particle size
distributions for contrasting soil suspensions using flow field-
flow fractionation. Environ Chem 3:184–191
Guo L, Wen L-S, Tang D, Santschi PH (2000) Re-examination of
cross-flow ultrafiltration for sampling marine colloids: evidence
from molecular probes. Mar Chem 69:75–90
Guzman KAD, Finnegan MP, Banfield JF (2006) Influence of surface
potential on aggregation and transport of titania nanoparticles.
Environ Sci Technol 40:7688–7693
Hall GEM (1998) Relative contamination levels observed in different
types of bottles used to collect water samples. Explorer 101:1–7
Handy RD, von der Kammer F, Lead JR, Hassello
¨
v M, Owen R,
Crane M (2008) The ecotoxicology and chemistry of manufac-
tured nanoparticles. Ecotoxicology 17:287–314
Hansen SF, Larsen BH, Olsen SI, Baun A (2007) Categorization
framework to aid hazard identification of nanomaterials. Nano-
toxicology 1:243–250
Hassello
¨
v M, Lyven B, Haraldsson C, Sirinawin W (1999) Determi-
nation of continuous size and trace element distribution of
colloidal material in natural water by on-line coupling of flow
field-flow fractionation with ICPMS. Anal Chem 71:3497–3502
Hassello
¨
v M, Lyven B, Bengtsson H, Jansen R, Turner DR, Beckett R
(2001) Particle size distributions of clay-rich sediments and pure
clay minerals: a comparison of grain size analysis with
sedimentation field-flow fractionation. Aquat Geochem 7:
155–171
Hassello
¨
v M, von der Kammer F, Beckett R (2007) Characterisation
of aquatic colloids and macromolecules by field-flow fraction-
ation. In: Wilkinson KJ, Lead JR (eds) Environmental colloids
and particles: behaviour, structure and characterization. John
Wiley and Sons, Chichester, pp 223–276
Hochella MF, Madden AS (2005) Earth’s nano-compartment for toxic
metals. Elements 1:199–203
Holt MS, Fox K, Griebach E, Johnsen S, Kinnunen J, Lecloux A,
Murray-Smith R, Peterson DR, Schro
¨
der R, Silvani M, ten Berge
WFJ, Toy RJ, Feijtel TCM (2000) Monitoring, modelling and
environmental exposure assessment of industrial chemicals in
the aquatic environment. Chemosphere 41:1799–1808
Hunter KA, Liss PS (1979) The surface charge of suspended particles
in estuarine and coastal water. Nature 282:823–825
Hyung H, Fortner JD, Hughes JB, Kim JH (2007) Natural organic
matter stabilizes carbon nanotubes in the aqueous phase. Environ
Sci Technol 41:179–184
Isaacson CW, Usenko CY, Tanguay RL, Field JA (2007) Quantifi-
cation of fullerenes by LC/ESI-MS and its application to in vivo
toxicity assays. Anal Chem 79:9091–9097
Jackson BP, Ranville JF, Bertsch PM, Sowder A (2005) Character-
ization of colloidal and humic-bound Ni and U in the
‘dissolved’’ fraction of contaminated sediment extracts. Environ
Sci Technol 39:2478–2485
Jiang L, Gao L, Sun J (2003) Production of aqueous colloidal
dispersions of carbon nanotubes. J Colloid Interface Sci 260:
89–94
Kim JI, Walther C (2007) Laser induced breakdown detection
(LIBD). In: Wilkinson KJ, Lead JR (eds) Environmental colloids
and particles: behaviour, structure and characterization. IUPAC
series on analytical and physical chemistry of environmental
systems. John Wiley and Sons, Chichester, pp 555–612
Larsson J, Gustafsson O, Ingri J (2002) Evaluation and optimization
of two complementary cross-flow ultrafiltration systems toward
isolation of coastal surface water colloids. Environ Sci Technol
36:2236–2241
Lead JR, Wilkinson KJ (2006) Aquatic colloids and nanoparticles:
current knowledge and future trends. Environ Chem 3:159–171
Lead JR, Wilkinson KJ, Balnois E, Cutak BJ, Larive CK, Assemi S,
Beckett R (2000a) Diffusion coefficients and polydispersities of
the Suwannee River fulvic acid: comparison of fluorescence
correlation spectroscopy, pulsed-field gradient nuclear magnetic
resonance, and flow field-flow fractionation. Environ Sci Tech-
nol 34:3508–3513
Lead JR, Wilkinson KJ, Starchev K, Canonica S, Buffle J (2000b)
Determination of diffusion coefficients of humic substances by
fluorescence correlation spectroscopy: role of solution condi-
tions. Environ Sci Technol 34:1365–1369
Lead JR, Muirhead D, Gibson CT (2005) Characterization of
freshwater natural aquatic colloids by atomic force microscopy
(AFM). Environ Sci Technol 39:6930–6936
Ledin A, Karlsson S, Duker A, Allard B (1994) Measurements in situ
of concentration and size distribution of colloidal matter in deep
groundwaters by photon-correlation spectroscopy. Water Res
28:1539–1545
Link S, El-Sayed MA (1999) Spectral properties and relaxation
dynamics of surface plasmon electronic oscillations in gold and
silver nanodots and nanorods. J Phys Chem B 103:8410–8426
Liu R, Lead JR (2006) Partial validation of cross flow ultrafiltration
by atomic force microscopy. Anal Chem 78:8105–8112
Madden AS, Hochella MF (2005) A test of geochemical reactivity as
a function of mineral size: manganese oxidation promoted by
hematite nanoparticles. Geochim Cosmochim Acta 69:389–398
Mavrocordatos D, Pronk W, Boller M (2004) Analysis of environ-
mental particles by atomic force microscopy, scanning and
transmission electron microscopy. Water Sci Technol 50:9–18
Mavrocordatos D, Perret D, Leppard GG (2007) Strategies and
advances in the characterization of environmental colloids by
electron microscopy. In: Wilkinson KJ, Lead JR (eds) Environ-
mental colloids and particles: behaviour, structure and
characterization. IUPAC series on analytical and physical
chemistry of environmental systems. John Wiley and Sons,
Chichester, pp 345–404
360 M. Hassello
¨
v et al.
123
Maynard AD, Aitken RJ, Butz T, Colvin V, Donaldson K, Oberdo
¨
r-
ster G, Philbert MA, Ryan J, Seaton A, Stone V, Tinkle SS, Tran
L, Walker NJ and Warheit DB (2006) Safe handling of
nanotechnology. Nature 444:267–269
Morrison MA, Benoit G (2001) Filtration artifacts caused by
overloading membrane filters. Environ Sci Technol 35:
3774–3779
Murr LE, Esquivel EV, Bang JJ, de la Rosa G, Gardea-Torresdey JL
(2004) Chemistry and nanoparticulate compositions of a 10,000
year-old ice core melt water. Water Res 38:4282–4296
Nowack B, Bucheli TD (2007) Occurrence, behavior and effects of
nanoparticles in the environment. Environ Pollut 150:5–22
Pal S, Tak YK, Song JM (2007) Does the antibacterial activity of
silver nanoparticles depend on the shape of the nanoparticle? A
study of the gram-negative bacterium Escherichia coli. Appl
Environ Microbiol 73:1712–1720
Perminova IV, Frimmel FH, Kudryavtsev AV, Kulikova NA, Abbt-
Braun G, Hesse S, Petrosyan VS (2003) Molecular weight
characteristics of humic substances from different environments
as determined by size exclusion chromatography and their
statistical evaluation. Environ Sci Technol 37:2477–2485
Ranville JF, Chittleborough DJ, Doss F, Harris T, Morrison R,
Beckett R (1999) Development of sedimentation field-flow
fractionation-inductively coupled plasma-mass spectrometry for
the characterization of environmental colloids. Anal Chim Acta
381:315–329
Roco MC (2005) International perspective on government nanotech-
nology funding in 2005. J Nanopart Res 7:707–712
Santos MC, Nobrega JA (2006) Slurry nebulization in plasmas for
analysis of inorganic materials. Appl Spectroscop Rev 41:
427–448
SCENIHR (2005) Opinion on the appropriateness of existing methodol-
ogies to assess the potential risks associated with engineered and
adventitious products of nanotechnology. Scientific Committee on
Emerging and Newly Identified Health Risks, European Commis-
sion. http://ec.europa.eu/health/ph_risk/committees/04_scenihr/
docs/scenihr_o_003b.pdf
Schimpf M, Caldwell K, Giddings JC (eds) (2000) Field-flow
fractionation handbook. John Wiley & Sons Inc., New York,
p 616
Schurtenberger P, Newman ME (1993) Characterization of biological
and environmental particles using static and dynamic light
scattering. In: Buffle J, van Leeuwen HP (eds) Environmental
particles. Lewis Publishers, Boca Raton, Florida, pp 37–115
Smith CJ, Shaw BJ, Handy RD (2007) Toxicity of single walled
carbon nanotubes to rainbow trout (Oncorhynchus mykiss):
respiratory toxicity, organ pathologies, and other physiological
effects. Aquat Toxicol 82:94–109
Suzuki Y, Kelly SD, Kemner KM, Banfield JF (2002) Radionuclide
contamination—nanometre-size products of uranium bioreduc-
tion. Nature 419:134–134
Taylor R (2006) Addition reactions of fullerenes. C R Chimie 9:
982–1000
Taylor R, Parsons JP, Avent AG, Rannard SP, Dennis TJ, Hare JP,
Kroto HW and Walton DRM (1991) Degradation of C60 by
light. Nature 351
Thiberge S, Nechushtan A, Sprinzak D, Gileadi O, Behar V, Zik O,
Chowers Y, Michaeli S, Schlessinger J, Moses E (2004)
Scanning electron microscopy of cells and tissues under fully
hydrated conditions. Proc Natl Acad Sci USA 101:3346–3351
Treubig JM, Brown PR (2002) Analysis of C60 and C70 fullerenes
using high-performance liquid chromatography–Fourier trans-
form infrared spectroscopy. J Chromatogr A 960:135–142
Viguie JR, Sukmanowski J, Nolting B, Royer FX (2007) Study of
agglomeration of alumina nanoparticles by atomic force micros-
copy (AFM) and photon correlation spectroscopy (PCS).
Colloids Surf Physicochem Eng Aspects 302:269–275
Vogl J, Heumann KG (1997) Determination of heavy metal
complexes with humic substances by HPLC/ICP-MS coupling
using on-line isotope dilution technique. Fresenius J Anal Chem
359:438–441
von der Kammer F (2005) Characterization of environmental colloids
applying field-flow fractionation—multi detection analysis with
emphasis on light scattering techniques. Hamburg University of
Technology, Hamburg, p 254
von der Kammer F, Baborowski M, Friese K (2005a) Application of
HPLC fluorescence detector as a nephelometric turbidity detec-
tor following field-flow fractionation to analyse size distributions
of environmental colloids. J Chromatogr A 1100:81–89
von der Kammer F, Baborowski M, Friese K (2005b) Field-flow
fractionation coupled to multi-angle laser light scattering
detectors: applicability and analytical benefits for the analysis
of environmental colloids. Anal Chim Acta 552:166–174
Waychunas GA, Kim CS, Banfield JF (2005) Nanoparticulate iron
oxide minerals in soils and sediments: unique properties and
contaminant scavenging mechanisms. J Nanopart Res 7:409–433
Wigginton NS, Haus KL, Hochella MF (2007) Aquatic environmental
nanoparticles. J Environ Monit 9:1306–1316
Williams A, Varela E, Meehan E, Tribe K (2002) Characterisation of
nanoparticulate systems by hydrodynamic chromatography. Int J
Pharm 242:295–299
Wyatt PJ (1998) Submicrometer particle sizing by multiangle light
scattering following fractionation. J Colloid Interface Sci 197:9–20
Yu WW, Qu L, Guo W, Peng X (2003) Experimental determination
of the extinction coefficient of CdTe, CdSe, and CdS nanocrys-
tals. Chem Mater 15:2854–2860
Nanoparticle analysis and characterization methods 361
123
... To obtain additional chemical information, UltraViolet-Visible (UV-Vis) spectrometry proves relevant, particularly in environmental studies to monitor the chromophoric groups present in natural organic matter such as humic substances. In particular, UV-Vis spectrometry can be used as a concentration detector with defined and constant operating conditions [19]. A UV-Vis Diode Array Detector (DAD) is also interesting to use for speciation by accessing the spectrum, and therefore molecular information [18,20,21]. ...
... Beforehand, the proportionality of the response of the organic carbon to its concentration in the water suspensions considered was verified as described by Harguindeguy et al. [33]. Indeed, a constant molar extinction coefficient at the selected wavelength is a prerequisite for using UV-Vis as a concentration detector [19]. A UV-Vis DAD (1260 Infinity series, Agilent Technology, Tokyo, Japan) was used to provide organic matter spectra. ...
Article
Full-text available
Copper (Cu) has been used to treat vines for a long time, which has led to its accumulation in vineyard soils. In the present work, the mobilization of copper from these soils and its transport, and diffusion outside the plots by drain water were investigated. For this, the distribution of copper between the dissolved and colloidal phases, and within the colloidal phase, of these waters was determined using an investigation strategy based on the coupling between a size separation technique, asymmetric flow field-flow fractionation, and several detectors. First, the total copper concentrations in water from different drains were monitored over a period of 2 years: Cu was mainly found in the fraction of < 450 nm. Then, the distribution of copper on the size continuum was more closely studied in water from one of the drains, sampled over a winter period. Between 45 and 75% of Cu was found in the 2–450 nm colloidal fraction. The <450 nm colloidal phase of the drain waters was found to be mainly composed of humic acids (~15 to 60 mg L−1) and clay-rich particles (~100 to 650 mg (Al) L−1). These particles also contained (hydr)oxides of iron and manganese. The concentrations of Fe and Mn were approximately 100 to 200 times lower than those of Al. The majority of humic acids had an apparent molar mass of ≤ 10 kDa. They were distributed along the size continuum: (i) in a population with an average size of ~20 nm, probably consisting of supramolecular entities, and (ii) associated with clay-rich particles with a size of ~120–200 nm. Copper was found to be complexed with humic acids and associated with clays via clay-humic complexes. Copper mobilization from the soil to the water and its transport to the drain water appeared governed by the soil humidity level and the rainfall.
... These findings provide important details on the relative quantities of different elements in the PLA samples that have been implanted with enzymes [13]. The information can be used to determine the components' possible effects on the environment or hazards, analyse the composition of the samples [14], and direct future studies or applications employing PLA materials with embedded enzymes [15]. ...
Article
Enzymes like lipases are essential for catalyzing important reactions in the fields of biotechnology and industrial operations. For quality control and safety evaluations, however, the quantification of these contaminants is crucial because the presence of heavy metals in enzyme preparations can negatively affect their activity and stability. Inductively Coupled Plasma Optical Emission Spectroscopy (ICP-OES), a key component of this investigation, is used to measure the levels of heavy metals in lipase enzyme preparations embedded in polylactic acid (PLA). The lipase enzyme being studied comes from Candida rugosa, which is well-known for its considerable industrial applications. According to preliminary findings, the lipase enzyme is successfully protected from external heavy metal pollutants by the encapsulating technique, maintaining its enzymatic activity and stability. The lipase-PLA composite had only minimal quantities of heavy metals, according to the ICP-OES study. The research's conclusions show great potential for the creation of reliable and contaminant-free lipase enzyme preparations, which will increase their suitability for usage in a variety of industrial processes and encourage the use of biodegradable polymers, supporting environmental sustainability. By highlighting the significance of quality control and safety assessment through the determination of heavy metal concentrations, this research contributes to the larger field of enzyme biotechnology. It also highlights the potential of Candida Rugosa lipase implanted in PLA matrices for eco-friendly and sustainable biocatalytic applications in sectors ranging from food and pharmaceuticals to biofuel generation and other biodegradable polymers.
... This technique measures the intensity of light scattering by a solution or suspension using a fluorescence spectrophotometer. 36,37 For NPs in suspension, Mie scattering is expected to be dominant. 38 This type of scattering is caused by relatively large particles and will therefore contribute only to the nephelometry signal when samples contain intact NPs. ...
Article
Full-text available
Analytical methods for the assessment of drug-delivery systems (DDSs) are commonly suitable for characterizing individual DDS properties, but do not allow determination of several properties simultaneously. A comprehensive online two-dimensional liquid chromatography (LC × LC) system was developed that is aimed to be capable of characterizing both nanoparticle size and encapsulated cargo over the particle size distribution of a DDS by using one integrated method. Polymeric nanoparticles (NPs) with encapsulated hydrophobic dyes were used as model DDSs. Hydrodynamic chromatography (HDC) was used in the first dimension to separate the intact NPs and to determine the particle size distribution. Fractions from the first dimension were taken comprehensively and disassembled online by the addition of an organic solvent, thereby releasing the encapsulated cargo. Reversed-phase liquid chromatography (RPLC) was used as a second dimension to separate the released dyes. Conditions were optimized to ensure the complete disassembly of the NPs and the dissolution of the dyes during the solvent modulation step. Subsequently, stationary-phase-assisted modulation (SPAM) was applied for trapping and preconcentration of the analytes, thereby minimizing the risk of analyte precipitation or breakthrough. The developed HDC × RPLC method allows for the characterization of encapsulated cargo as a function of intact nanoparticle size and shows potential for the analysis of API stability.
Article
Full-text available
Commercial applications of nanotechnology in the food industry are rapidly increasing. Accordingly, there is a simultaneous increase in the amount and diversity of nanowaste, which arise as byproducts in the production, use, disposal, or recycling processes of nanomaterials utilized in the food industry. The potential risks of this nanowaste to human health and the environment are alarming. It is of crucial significance to establish analytical methods and monitoring systems for nanowaste to ensure food safety. This review provides comprehensive information on nanowaste in foods as well as comparative material on existing and new analytical methods for the detection of nanowaste. The article is specifically focused on nanowaste in food systems. Moreover, the current techniques, challenges as well as potential use of new and progressive methods are underlined, further highlighting advances in technology, collaborative efforts, as well as future perspectives for effective nanowaste detection and tracking. Such detection and tracking of nanowaste are required in order to effectively manage this type ofwasted in foods. Although there are devices that utilize spectroscopy, spectrometry, microscopy/imaging, chromatography, separation/fractionation, light scattering, diffraction, optical, adsorption, diffusion, and centrifugation methods for this purpose, there are challenges to be overcome in relation to nanowaste as well as food matrix and method characteristics. New technologies such as radio‐frequency identification, Internet of things, blockchain, data analytics, and machine learning are promising. However, the cooperation of international organizations, food sector, research, and political organizations is needed for effectively managing nanowaste. Future research efforts should be focused on addressing knowledge gaps and potential strategies for optimizing nanowaste detection and tracking processes.
Chapter
The degradation by biological processes of a large majority of organic volatile contaminants in wastewater is limited. This is so, largely because the reaction is inhibited at concentrations higher than 50 mg.L−1, thus, requiring costly physical or physiochemical pre-treatment(s). Electrochemistry currently plays an important role in a vast number of fundamental studies and applied areas in the treatment of pharmaceutical effluents. Pharmaceutical effluents are wastewater generated through various pharmaceutical processes such as the manufacturing of active pharmaceutical ingredients and product formulation. This wastewater is characterized by high concentrations of organic matter, catalysts such as TiO2/g-C3N4 and cathodic WO3/W nanocatalysts, and toxic pollutants that include Benzene, ethylbenzene, and Toluene). In recent times, electrochemical treatment has been attracting researchers as an emerging technology used for the removal of organic and inorganic impurities from water and wastewater. Many scientists are attempting to use these methods for the treatment of wastewater owing to the many advantages associated with electrochemistry compared to their biological or biotechnological counterparts. Although several review papers are available regarding the application of electrochemical methods for the environmental clean-up of pharmaceutical wastewater, research focussing on the development of efficient technologies coupled with the use of electrochemical and photocatalytic nanocatalysts continues. This chapter presents a detailed review of methods for the electrochemical degradation of volatile organic compounds in wastewater catalysed by nanocatalysts with an emphasis on pharmaceutical wastewater or effluents. An extensive review of the mechanism and application of these nanocatalysts-promoted electrochemical processes for degradation, mineralization, and detoxification of different organic pollutants present in industrial pharmaceutical wastewater will be reported. Electrochemical methods for the destruction of organic pollutants in wastewater is becoming very attractive in the industry because it is a powerful and promising clean method with undisputable environmental friendliness. Furthermore, the electrochemical treatment methods can be set up easily and show high efficiency. Analytical challenges encountered in the physical and electrochemical characterisation of some of these nanocatalysts will be briefly discussed. The reason for an accurate analytical characterisation is to check the composition, structure, dispersion on, and active surface area of the nanocatalysts, as well as the electrochemical systems. This guarantees the quality of the electrochemical data obtained. Finally, electrochemical and differential electrochemical mass spectrometry measurements used in specific cases to study and evaluate the influence of the electrocatalyst structure on its electroactivity will be discussed. This type of study allows for the investigation of the effect of the composition in terms of foreign metal atoms and atomic content of nano-metals-based catalysts (e.g. ZnO, TiO2, or CuO) towards the electro-oxidation of the organic pollutants. This chapter concludes by presenting the summary of the principles, advantages, and limitations, as well as the performances and some of the successful applications of the nanocatalysed electrochemical degradation of volatile organic compounds in wastewater using pharmaceutical effluents as an example.
Chapter
Full-text available
Software and Programming Tools in Pharmaceutical Research is a detailed primer on the use for computer programs in the design and development of new drugs. Chapters offer information about different programs and computational techniques in pharmacology. The book will help readers to harness computer technologies in pharmaceutical investigations. Readers will also appreciate the pivotal role that software applications and programming tools play in revolutionizing the pharmaceutical industry. The book includes nine structured chapters, each addressing a critical aspect of pharmaceutical research and software utilization. From an introduction to pharmaceutical informatics and computational chemistry to advanced topics like molecular modeling, data mining, and high-throughput screening, this book covers a wide range of topics. Key Features: - Practical Insights: Presents practical knowledge on how to effectively utilize software tools in pharmaceutical research. - Interdisciplinary Approach: Bridges the gap between pharmaceutical science and computer science - Cutting-Edge Topics: Covers the latest advancements in computational drug development, including data analysis and visualization techniques, drug repurposing, pharmacokinetic modelling and screening. - Recommendations for Tools: Includes informative tables for software tools - Referenced content: Includes scientific references for advanced readers The book is an ideal primer for students and educators in pharmaceutical science and computational biology, providing a comprehensive foundation for this rapidly evolving field. It is also an essential resource for pharmaceutical researchers, scientists, and professionals looking to enhance their understanding of software tools and programming in drug development.
Article
mTitanium dioxide nanoparticles (TiO2 NPs) have become a focal point of research due to their widespread daily use and diverse synthesis methods, including physical, chemical, and environmentally sustainable approaches. These nanoparticles possess unique attributes such as size, shape, and surface functionality, making them particularly intriguing for applications in the biomedical field. The continuous exploration of TiO2 NPs is driven by the quest to enhance their multifunctionality, aiming to create next-generation products with superior performance. Recent research efforts have specifically focused on understanding the anatase and rutile phases of TiO2 NPs and evaluating their potential in various domains, including photocatalytic processes, antibacterial properties, antioxidant effects, and nanohybrid applications. The hypothesis guiding this research is that by exploring different synthesis methods, particularly chemical and environmentally friendly approaches, and incorporating doping and co-doping techniques, the properties of TiO2 NPs can be significantly improved for diverse applications. The study employs a comprehensive approach, investigating the effects of nanoparticle size, shape, dose, and exposure time on performance. The synthesis methods considered encompass both conventional chemical processes and environmentally friendly alternatives, with a focus on how doping and co-doping can enhance the properties of TiO2 NPs. The research unveils valuable insights into the distinct phases of TiO2 NPs and their potential across various applications. It sheds light on the improved properties achieved through doping and co-doping, showcasing advancements in photocatalytic processes, antibacterial efficacy, antioxidant capabilities, and nanohybrid applications. The study concludes by emphasizing regulatory aspects and offering suggestions for product enhancement. It provides recommendations for the reliable application of TiO2 NPs, addressing a comprehensive spectrum of critical aspects in TiO2 NP research and application. Overall, this research contributes to the evolving landscape of TiO2 NP utilization, offering valuable insights for the development of innovative and high-performance products.
Article
Full-text available
Stable homogeneous dispersions of carbon nanotubes (CNTs) have been prepared by using sodium dodecyl sulfate (SDS) as dispersing agent. To our knowledge, it is the first report to quantitatively characterize colloidal stability of the dispersions by UV-vis spectrophometric measurements. When the sediment time reaches 500 h, the supernatant CNT concentration drops as much as 50% for the bare CNT suspension, compared to 15% with the addition of SDS. Furthermore, after 150 h, no precipitation is found for CNT/SDS dispersions, exhibiting an extreme stability. Zeta potential, auger electron microscopy, and FTIR analysis are employed to investigate the adsorption mechanism in detail. It has been concluded that the surfactant containing a single straight-chain hydrophobic segment and a terminal hydrophilic segment can modify the CNTs-suspending medium interface and prevent aggregation over long periods. The morphology of the CNT dispersions is observed with optical microscopy. An intermediate domain of homogeneously dispersed nanotubes exhibits an optimum at 0.5 wt% CNTs and 2.0 wt% SDS.
Chapter
IntroductionSampling of Environmental Colloids for Electron Microscopic InvestigationSpecimen PreparationMorphometric Analysis of Particles and ColloidsElement Analysis of Particles and ColloidsApplications of EM and AEM for the Understanding of Physicochemical Pathways in the EnvironmentConclusion List of abbreviationsReferences
Article
The field of nanoparticle research has drawn much attention in the past decade as a result of the search for new materials. Size confinement results in new electronic and optical properties, possibly suitable for many electronic and optoelectronic applications. A characteristic feature of noble metal nanoparticles is the strong color of their colloidal solutions, which is caused by the surface plasmon absorption. This article describes our studies of the properties of the surface plasmon absorption in metal nanoparticles that range in size between 10 and 100 nm. The effects of size, shape, and composition on the plasmon absorption maximum and its bandwidth are discussed. Furthermore, the optical response of the surface plasmon absorption due to excitation with femtosecond laser pulses allowed us to follow the electron dynamics (electron−electron and electron−phonon scattering) in these metal nanoparticles. It is found that the electron−phonon relaxation processes in nanoparticles, which are smaller than the electron mean free path, are independent of their size or shape. Intense laser heating of the electrons in these particles is also found to cause a shape transformation (photoisomerization of the rods into spheres or fragmentation), which depends on the laser pulse energy and pulse width.
Article
Semiconductor quantum dots (QDs) are nanometer-sized crystals with unique photochemical and photophysical properties that are not available from either isolated molecules or bulk solids. In comparison with organic dyes and fluorescent proteins, these quantum-confined nanoparticles are brighter, more stable against photobleaching, and can be excited for multicolor emission with a single light source. Recent advances have shown that nanometer-sized semiconductor particles can be covalently linked with biorecognition molecules such as peptides, antibodies, nucleic acids, or small-molecule ligands for use as biological labels. High-quality QDs are also well suited for optical encoding and multiplexing applications due to their broad excitation profiles and narrow/symmetric emission spectra. In this article, we discuss recent developments in QD synthesis and bioconjugation, their applications in molecular and cellular imaging, as well as promising directions for future research.