You are on page 1of 7

Journal of Colloid and Interface Science 222, 83–89 (2000)

doi:10.1006/jcis.1999.6599, available online at http://www.idealibrary.com on

Structure of Hydrous Ferric Oxide Aggregates


Bill Lo and T. David Waite1
School of Civil and Environmental Engineering, The University of New South Wales, Sydney, New South Wales 2052, Australia

Received June 4, 1999; accepted October 19, 1999

size was observed to decrease during the first 30 min of hydrol-


The small angle light scattering behavior of hydrous ferric ox- ysis/nucleation with the final mean size centered around 10 nm.
ide flocs is examined here and found to provide useful insights into Highly linear structures were observed at low OH/Fe ratios (ap-
the nature of the aggregates formed despite the large size of these
parently as a result of long-range magnetic dipolar interactions)
aggregates at later times. The flocs appear to exhibit fractal prop-
erties over a significant size range though the aggregates appear
but became more branched (exhibiting fractal dimensions in the
to be easily disrupted through mixing effects resulting in breakup 1.7 to 2.0 range) at higher OH/Fe ratios.
and/or restructuring to denser assemblages. Background electrolyte Once flocculation of these “precursor” colloids is induced (for
concentrations also have some impact on floc structure but mixing example, by reduction in surface charge through pH increase),
effects and apparent destabilization by ferric ions limit the effect large aggregates are rapidly formed. These large aggregates are
of added electrolytes on the stability and structure of ferric oxyhy- typical of those formed in water and wastewater treatment where
droxides. Similar estimates of fractal dimensions of these hydrous the large size is a factor in their ease of sedimentation or their
ferric oxide flocs are obtained both by static light scattering anal- removal by filtration. The structure of these aggregates remains
ysis and by a cluster mass scaling approach. The choice of density of interest since it will be a determinant (together with cluster
distribution cutoff function has some impact on derived size and size) of their permeability—a factor of obvious importance in
structure parameters and further refinement in this area is needed.
size separation processes.
°
C 2000 Academic Press
Key Words: iron oxide; ferrihydrite; fractal; aggregate; light scat-
Given the large (relative) size of these aggregates, small angle
tering; structure. light scattering represents one possible means of probing their
structure. The light scattering behavior of aggregating amor-
phous iron oxides under various system conditions (electrolyte
INTRODUCTION concentration, mixing regime) is reported here and, based on the
results presented, conclusions are drawn on their structure as a
Iron oxides occur widely in nature and are used extensively function of system conditions including background electrolyte
in water and wastewater treatment because of their ability to ad- concentration and mixing regime.
sorb contaminants. The so-called “amorphous” iron oxide (also
known as hydrous ferric oxide and ferrihydrite) is of particular THEORETICAL
interest as it is the form that typically occurs initially in natural
aqueous environments prior to aging to more crystalline forms In a scattering experiment, a beam of light (or X rays or neu-
and is the form that occurs in treatment systems when ferric trons) is directed onto a sample and the scattered intensity I
salts are hydrolyzed under neutral to alkaline conditions. The (in photon counts) is measured as a function of wave number
crystallinity (as determined by X-ray diffractometry) of amor- q = (4π n/λ) sin(θ/2) where θ is the scattering angle, λ is the
phous iron oxide may vary from the highly disordered 2-line wavelength of the incident beam, and n is the refractive index of
ferrihydrite to the more ordered 6-line form (1). the medium. If N (M) denotes the number of clusters of mass M
Considerable attention has been given to the structure of col- in the scattering volume and if interactions between the clusters
loidal amorphous iron oxide in its nonflocculated form through are negligible, the total scattering intensity is given by (4)
application of small angle X-ray scattering studies (2, 3). Thus,
X
Tchoubar et al. (2) report the formation of colloidal aggregates I (q) = N (M)I M (q), [1]
whose size (at 400 s after commencement of hydrolysis) ranged M
from 10 nm at low OH/Fe ratios to around 700 nm at high OH/Fe
ratios (but still below the flocculation threshold). The colloid where I M (q) is the scattered intensity from clusters of mass M.
The choice of the scattering angle sets a length scale, given
by q −1 , on which the clusters are probed. This length scale must
1 To whom correspondence should be addressed: Fax: +61-2-9385 6139. be compared with a suitably defined length scale of a cluster of
E-mail: d.waite@unsw.edu.au. mass M. Choosing the radius of gyration Rg , one may consider
83 0021-9797/00 $35.00
Copyright °C 2000 by Academic Press
All rights of reproduction in any form reserved.
84 LO AND WAITE

the two limits q Rg ¿ 1 and q Rg À 1. In the first case, the cluster Meriani (9) have also argued that the Gaussian cutoff function
appears to the light probe essentially as a point particle and thus provides a better description of light scattering by finite-sized
scatters isotropically, independent of q. Therefore, I M (q) ∼ M 2 fractal aggregates. Expressions for ξ 2 and S(q) for the Gaussian
(4). In the second case, q Rg À 1, one may decompose the clus- cutoff function are given below
ter of radius Rg into smaller subunits (termed “blobs” by Lin
et al. (4)) of size q −1 . As long as these subunits contain many 4Rg2
colloidal particles and these particles exibit a fractal structure ξ2 = [4]
Df
within the subunit, the subunit mass will scale as m su ∼ (qa)−Df , µ ¶
where “a” denotes the radius of an individual colloidal particle −(q Rg )2 /Df ∗ 3 − Df 3 (q Rg )2
S(q) = e 1 F1 , ; , [5]
and Df is the fractal dimension. The phase differences between 2 2 Df
the scattered fields from particles within a single subunit are al-
ways less than one. Thus, a single subunit scatters coherently, where 1 F1 is the hypergeometric function.
and the scattered intensity of the subunit will be proportional It should be noted that the assumption that the phase difference
to m 2su . By contrast, the phase differences between the scattered between scattered waves depends only upon their location in the
fields from different subunits are always greater than one. Thus, cluster and is independent of any material property of the cluster
the phases of the electric fields scattered from different subunits (the Rayleigh-Gans-Debye approximation) is valid as long as
add incoherently and the total intensity will be the sum of the
intensities scattered from each subunit. The number of subunits |m − 1| ¿ 1 [6a]
in a cluster of mass M is M/m su ∼ M(qa) Df . Therefore, the 4πa
total scattered intensity is I M (q) ∼ (M/m su )m 2su ∼ M(qa)−Df or |m − 1| ¿ 1, [6b]
λ
I M (q) ∼ M 2 (q Rg )−Df (4).
Since the factor M 2 appears in both limiting cases, it is con- where m is the refractive index of the particle relative to the sus-
venient to write I M (q) = M 2 SM (q), where S M (q) denotes the pending medium (10). It is this condition that requires |m − 1|
“structure factor” of clusters of mass M (4). The total scattered to become smaller and smaller as the RGD theory is applied to
intensity (from Eq. [1]) is therefore given by progressively larger particles.
X
I (q) = N (M)M 2 S M (q), [2]
M EXPERIMENTAL

which is a measure of both the cluster mass distribution N (M) Ferrihydrite Preparation and Aggregation
and the structure of the clusters.
The structure factor reflects, in q space, the distribution of The amorphous iron oxide, ferrihydrite, was prepared by
scattering material in real space; S M (q) is the Fourier trans- adding FeCl3 from a stock solution of 0.5 wt% to a 500-mL
form of g(r ), the density correlation function. For a three- mixture of 1 mM NaHCO3 and either 1 or 100 mM NaCl. Homo-
dimensional, self-similar object with fractal dimension Df , this geneity was achieved by gentle swirling. Solutions of 10 mg/L
function scales as g(r ) ∼ r Df −3 . For colloidal aggregates, how- FeCl3 were typically prepared which, at the pH induced by the
ever, self-similar scaling occurs only over a limited range of carbonate/CO2 buffer (pH 7.8), rapidly hydrolyzed then nucle-
length scales. The upper limit for the scaling form of g(r ) oc- ated to colloidal amorphous iron oxide. Aliquots (8 mL) of these
curs when r reaches the radius of gyration, Rg . The lower limit suspensions were immediately transferred to the small cell of the
of the scaling is given by the primary particle radius, a. The ex- Malvern Mastersizer where the amorphous iron oxides contin-
istence of the upper limit introduces a cutoff function h(r/Rg ) ued to aggregate. This small cell either was not stirred or was
such that stirred at one of two stirrer speeds (“moderately” or “rapidly”)
using a magnetic stirring bar.
g(r ) ∼ r Df −3 h(r/Rg ), [3]
Aggregate Size and Structure Measurement
with h(x) = 1 for x ¿ 1 and h(x) → 0 for x À 1. A variety of Size distributions and structural information of amorphous
cutoff functions have been used (5–7) including exponential, iron oxide aggregates were determined as a function of aggre-
Gaussian, “overlapping sphere,” and stretched exponential ex- gation time using a Malvern Mastersizer/E which ascertains size
pressions. The exponential cutoff function h(r/ξ ) = e−(r/ξ ) is by analysis of forward scattered light. The Mastersizer consists
the simplest and has the correct asymptotic behavior but has no of a 5 mW He-Ne laser (632.8-nm wavelength) with 18-mm
physical basis. It has been extensively used though Sorensen and beam expansion and collimation. The direction of the polarized
co-workers (6, 7) have shown that this function is inappropriate laser is parallel to the detector axis (vertical). The particles pass
if the system exhibits some polydispersity and argue that the through the expanded and collimated laser beam in front of the
Gaussian cutoff function h(r/ξ ) = e−(r/ξ ) is a better choice. A
2
optic lens in focal plane of which are positioned 31 photosensi-
number of other authors including Jullien (8) and Yanwei and tive detectors. The Fourier optics can be used with 45-, 100-, or
STRUCTURE OF HYDROUS FERRIC OXIDE AGGREGATES 85

300-mm lenses which allow the collection of light scattered at


angles from 0.01 to 32.5◦ . In this work, the flocs were analyzed
using the 300-mm lens, enabling the collection of light scattered
from 0.03 to 6.25◦ .
Size distribution information was obtained using the software
supplied and involved calculation of the size distribution by an
iterative algorithm which matches the measured and calculated
energy distributions over the detectors. The fitting procedure is a
model-independent technique (i.e., it does not constrain the vol-
ume distribution to follow some common analytical expression
such as the log normal distribution function) and uses Fraun-
hoffer theory to develop a scattering pattern assuming a popu-
lation of equivalent spheres. Information on distribution size is
presented in this paper principally as volume average diameter
(VAD) where
X ÁX
VAD = (Vi di ) Vi , [7]
i i
FIG. 2. Conventional static light scattering plots (log I versus log q) for
where Vi is the relative volume in size class i with mean class amorphous iron oxides as a function of aggregation time.
diameter di .
Information on assemblage structure was obtained by mea- distributions obtained from the Malvern fits to the light scat-
suring the intensity of light scattered at all available detectors tering data are reasonably log normal in shape with relatively
and plotting log intensity versus log q. Information required to narrow, consistent spread in aggregate size over time.
compute the angle of each detector to the incident beam as well The log I versus log q light scattering plots shown in Fig. 2
as size and responsiveness of each detector (the so-called “magic reflect increasing power law behavior with aggregation time
numbers”) was supplied by Malvern Instruments. with at least one decade of linearity (10−4 nm−1 < q < 2 ×
10−3 nm−1 ) observed after about 6 min of aggregation. Sim-
RESULTS AND DISCUSSION ilar behavior, albeit over considerably longer aggregation times,
has previously been reported for silica particles (11).
With gentle mixing in the Malvern cell, the (equivalent spher- Given the large size of particles developed in these amor-
ical) size of amorphous iron oxide aggregates evolve over time phous iron oxide aggregation studies, the question of whether
from distributions exhibiting modal diameters in the low mi- the primary particles maintain independent scattering must be
crometer range at early times (within the first minute after addi- raised before interpreting these scattering plots further (i.e., is
tion to the Malvern cell) to modal diameters in excess of 100 µm the Rayleigh-Gans-Debye approximation valid). As discussed
after a 15-min aggregation time. As can be seen from Fig. 1, the earlier, the RGD approximation is valid when both |m − 1| ¿ 1
and 4πa/λ|m − 1| ¿ 1 are satisfied. Of prime importance here
is the nature of the primary scatterers. As discussed in the intro-
duction, Bottero and co-workers (2, 3) have shown that colloidal
ferrihydrite particles are initially formed which exhibit fractal di-
mensions in the 1.7–2.0 range and, as such, are relatively porous
and would be expected to exhibit low refractive indices (most
likely m < 1.1). The size of these aggregates (which may be
considered our primary scatterers at the wave vector (q) mag-
nitudes being used here) is unclear and, from Tchoubar et al.’s
work (2), could range from 10’s to 100’s of nanometres. Elec-
tron microscopy studies reported by Cornell and Schwertman
(1) suggest that these “core” aggregates are in the 50 to 300 nm
size range. Given that these primary particles are themselves
aggregates of considerable porosity, their refractive index will
be much closer to unity than that of crystalline iron oxides. For
example, at a refractive index of 1.1 and a primary scatterer size
(a) of (say) 300 nm, the RGD conditions may be considered
FIG. 1. Amorphous iron oxide size distributions obtained using the Malvern to be satisfactorily met. It should also be noted that early work
Mastersizer as a function of aggregation time. by Kerker (12) and more recent studies by Farias et al. (13)
86 LO AND WAITE

(average) radius of gyration of the clusters. In reality, of course,


a distribution in cluster sizes exists and, as might be expected,
improved fits of the scattering model to the data are obtained by
introducing some polydispersity in cluster size (as defined by
a log normal distribution in radius of gyration) into the fitting
process (see Fig. 3). As shown in Table 1, increasing “spread”
in Rg is observed to have only a small effect on the estimated
Df value.
Care should be taken in interpreting the derived Df values as a
definitive measure of the fractal dimension of the assemblage as
the choice of cutoff function clearly has an impact on the values
obtained by fitting of the scattering equation to the obtained
I versus q data. While we have adopted (for good reason) a
Gaussian cutoff function in all data analysis presented here, it
is clear that some shift in parameter values might be obtained
if alternate descriptions were adopted. For this reason, we will
consistently refer to Rg and Df values as “apparent” radii of
gyration and fractal dimensions, respectively.
Fits of the full light scattering equation to aggregates devel-
oped in the Malvern cell with moderate mixing are shown in
Fig. 2. A Gaussian cutoff function (assuming an average Rg
FIG. 3. Light scattering plot (at t = 6 min) showing Gaussian fit assuming value in all cases) has been used in all fits. The time evolution
both a single (average) Rg value and a log normal distribution of Rg values. in selected measured (Imax and VAD) and fitted (Rg and Df )
parameters is given in Table 2.
indicate that these conditions may be relaxed considerably, par- The results reported in Table 2 suggest an increase in “com-
ticularly for the case of low angles of scatter. Indeed, Heimenz pactness” of amorphous iron oxide aggregates over the course
and Rajagopalan (14) note that for a scattering angle of 10◦ , the of the aggregation process and may reflect the fragility of these
RGD approximation is good to within 10% for spheres with a aggregates and their susceptibility to breakup and/or restructur-
radius that is about 62, 37, and 25 times that of the wavelength ing on mixing. The effects of mixing appear to be even more
for refractive indices of 1.1, 1.2, and 1.3, respectively. Increasing distinct in the results shown in Fig. 4. In this case, flocs have
effects of multiple scattering and interaction between particles,
however, are to be expected as the aggregate size and compact-
TABLE 2
ness increase.
Measured and Derived Parameter Values for Amorphous Iron
As outlined under theoretical, the full light scattering data
Oxides as a Function of Aggregation Time (in Small Malvern Mea-
set can be used to obtain an estimate of the fractal dimension surement Cell with Moderate Mixing)a
(Df ) and the size cutoff value (ξ ) (or radius of gyration, Rg )
provided an appropriate expression for the density correlation Time VAD Imax Rg
cutoff function is used. The best fit result for t = 6 min data using (min) (µm) (arbitrary units) (µm) Df
a Gaussian cutoff function is shown in Fig. 3 and fit parameters
0 5.2 5.4 3.0 1.48
shown in Table 1. These results are obtained assuming a single 1 6.3 7.5 3.1 1.85
2 8.2 21.6 3.8 2.09
3 11.1 66.4 4.8 2.22
TABLE 1 4 14.9 89.1 6.7 2.27
Average Radius of Gyration (Rg ) and Apparent Fractal Dimen- 5 19.9 221.5 9.4 2.27
sions (D f ) Obtained Using a Gaussian Cutoff Function in the Light 6 27.2 654.3 12.4 2.33
Scattering Equation Used to Fit Log I Versus Log q Data Obtained 7 36.4 981.8 15.2 2.36
after Aggregation with Moderate Stirring for 6 mina 8 48.6 1782.0 20.0 2.39
9 60.5 2953.0 24.0 2.41
Rg 10 67.8 4338.9 27.3 2.45
(µm) Df 11 80.0 6523.2 32.1 2.46
12 92.9 9078.4 33.0 2.47
Single (average) Rg 12.35 2.33 13 111.7 14603.4 33.5 2.54
Distribution of Rg 12.35 ± 1σ 2.42 14 121.3 17258.0 33.9 2.56
15 128.3 19953.5 34.0 2.58
a Also shown is the effect of accounting for the presence of polydispersity

in cluster size by using a log normal distribution of Rg values (a log σ = 0.2 is a Radius of gyration and apparent fractal dimension values are obtained by

assumed). fitting the complete light scattering equation using a Gaussian cutoff function.
STRUCTURE OF HYDROUS FERRIC OXIDE AGGREGATES 87

TABLE 3
Effect of Background Electrolyte Concentration on Fractal Di-
mension of Resulting Aggregates under Conditions of No Mixing
and “Moderate” Mixing in the Small Malvern Measurement Cell

Df

[NaCl] Analysis time No mixing Moderate mixing

10 mM 15 min 2.07 2.58


100 min 2.29
100 mM 15 min 1.99 2.50
100 min 2.08

10 to 100 mM (the concentration range normally found to induce


colloid destabilization) did not have a large effect on aggrega-
tion rate, presumably because sufficient ferric ions are already
present in solution to induce rapid aggregation. As shown in
Table 3, a small effect of salt in inducing a more open struc-
ture is observed though the results are by no means as dramatic
as those in standard cases of reaction-limited versus diffusion-
limited aggregation where fractal dimensions of 1.7–1.8 are typi-
cal of DLA while fractal dimensions of 2.2–2.3 are more typical
of RLA. The obtained results are also complicated by the in-
crease in fractal dimension over the course of aggregation that
was described above.
While an observable difference in fractal dimensions of flocs
is observed under low salt compared to high salt conditions when
FIG. 4. Effect of mixing conditions on volume average diameter and ap- aggregates are formed with no stirring, the presence of mixing
parent fractal dimension of amorphous iron oxide flocs over time. in the cell leads to rapid formation of more compact aggregates
for which the Df , once in the 2.5–2.6 range, appears relatively
stable.
been gently added to the measurement cell and either the cell A number of additional insights may be gleaned from the in-
not stirred or the cell stirred via a magnetic stirring bar rotating formation presented in Table 2. For example, a strongly linear
either slowly (“moderate” stirring) or rapidly (“rapid” stirring). correlation is observed between the measured volume average
As can be seen from the floc size and structure results over time, diameter and the estimated radius of gyration (Fig. 5). Such a
mixing condition has a dramatic impact upon the size and appar- relationship is to be expected if the flocs maintain a uniform
ent structure of the resulting flocs. With no stirring, flocs grow shape as aggregation occurs. Additionally, at low q, where the
very slowly (<20 µm in 20 min) and light scattering intensity
is low. The apparent fractal dimension of these flocs eventually
approaches a relatively constant value of 2.1–2.2. With “moder-
ate” stirring, the VAD exhibits an exponential increase over time
to about 100 µm after only 10 min and the apparent fractal di-
mension increases from under 1.9 to approximately 2.5 in about
10 min. With more rapid stirring, the flocs commence growing
very quickly (as measured by the VAD) reaching almost 50 µm
in about 2 min. After this time, however, floc growth is curtailed
with a reduction in volume average diameter observed on con-
tinued stirring. The apparent fractal dimension is observed to
rapidly increase on this 2- to 3-min time scale and plateaus at
just over 2.5 after about 5 min of mixing.
Another factor that has a significant influence on aggregation
kinetics and resulting aggregate structure in many colloidal sys- FIG. 5. Radius of gyration (Rg ) as deduced from modeling of static light
tems is the background electrolyte concentration. In the case of scattering data for hydrous ferric oxide aggregates using Gaussian cutoff function
ferrihydrite, increasing electrolyte (NaCl) concentrations from versus volume average diameter (VAD) of the aggregates.
88 LO AND WAITE

scattered intensity reflects scattering from the complete aggre- cutoff function) or may indicate that the value obtained in the
gates and the scattered intensity reaches a q-invariant maximum, low q region using the mass scaling behavior (Mα Imax α RgDf ) is
Imax ∝ MN (M)M. Now for an aggregating system where we are an averaged value with small scale change indiscernible.
not creating or losing mass (assuming no sedimentation is oc- Results for low salt conditions (10 mM NaCl) are shown in
curring), we would expect N (M)M to be constant. If this is the Fig. 6a while those for a high salt concentration (100 mM NaCl)
case, then we would expect to see Imax ∝ M. For a fractal ag- are given in Fig. 6b. Satisfyingly, the fractal dimensions esti-
gregate for which M ∝ RgDf , Imax should increase according to mated from the slope of the linear portion of these plots (2.68
RgDf ; that is, a plot of log Imax versus log Rg should provide a and 2.61 for the 10 and 100 mM NaCl cases, respectively) are
measure of the fractal dimension. Such plots are shown in Fig. 6 not too dissimilar to the values derived from analysis of the static
and, apart from results at early times when the flocs are small light scattering data (I vs q plots) using the full scattering equa-
and still establishing their fractal character, do exhibit linear re- tion incorporating the Gaussian cutoff function (2.58 and 2.50
lationships between log Imax and log Rg over a significant size for the 10 and 100 mM NaCl cases, respectively). Even closer
range. The fact that the log Imax versus log Rg plots are linear and agreement could no doubt be obtained by incorporating some
not showing a gradual increase in Df with time (as found from degree of polydispersity into the radius of gyration term used in
the log I versus log q fits) may suggest that the change in Df val- fitting the full scattering equation as discussed earlier.
ues derived from the log I versus log q plots is a function of the
fitting process (a possibility, especially when using a Gaussian
CONCLUSIONS

Static light scattering appears to be a useful tool for investi-


gation of the structure of ferric oxyhydroxide flocs despite the
large size of aggregates formed. The finite size of ferric oxide
flocs examined necessitated the use of a density distribution cut-
off function in fitting the complete light scattering data with the
choice of function having some impact on size and structure
parameters determined.
The flocs appear to exhibit fractal properties over a signifi-
cant size range though this fractal structure appears to be easily
disrupted through mixing effects resulting in breakup and/or
restructuring to dense aggregates. Background electrolyte con-
centrations also have some impact on apparent floc structure but
mixing effects and apparent destabilization by ferric ions limit
the effect of added electrolytes on the stability and structure of
ferric oxyhydroxides. Similar estimates of fractal dimension of
hydrous ferric oxide flocs are obtained by analysis of both the
static light scattering (I vs q) data and the cluster mass scaling
(Imax α M vs Rg ) behavior.

ACKNOWLEDGMENTS

The CRC for Waste Management and Pollution Control is thanked for the
provision of a postgraduate research scholarship to Mr. Lo.

REFERENCES

1. Cornell, R. M., and Schwertmann, U., “The Iron Oxides–Structure,


Properties, Reaction, Occurrences and Uses.” VCH Publ., Weinheim,
1996.
2. Tchoubar, D., Bottero, J.-Y., Quienne, P., and Arnaud, M., Langmuir 7, 398
(1991).
FIG. 6. Log-log plot of maximum scattering intensity (Imax ) versus Rg for 3. Bottero, J.-Y., Manceau, A., Villieras, F., and Tchoubar, D., Langmuir 10,
the flocculation of amorphous iron oxides in the presence of (a) 10 mM NaCl 316 (1994).
and (b) 100 mM NaCl. Data points used in regression analysis shown as solid 4. Lin, M. Y., Klein, R., Lindsay, H. M., Weitz, D. A., Ball, R. C., and Meakin,
symbols. P., J. Colloid Interface Sci. 137, 263 (1990).
STRUCTURE OF HYDROUS FERRIC OXIDE AGGREGATES 89

5. Schmidt, P. W., in “The Fractal Approach to Heterogeneous Chemistry: Sur- 11. Cannell, D. S., and Aubert, C., in “On Growth and Form: Fractal and Non-
faces, Colloids, Polymers” (D. Avnir, Ed.), p. 67. Wiley, New York, 1989. Fractal Patterns in Physics” (H.E. Stanley and N. Ostrowsky, Eds.), p. 187.
6. Sorensen, C. M., Cai, J., and Lu, N., Langmuir 8, 2064 (1992). Martinus Nijhoff Publ., Dordrecht, 1986.
7. Sorensen, C. M., in “Surface and Colloid Chemistry” (K. S. Birdi, Ed.), 12. Kerker, M., “The Scattering of Light and Other Electromagnetic Radiation.”
Chap. 13, p. 533. CRC Press, Boca Raton, FL, 1997. Academic Press, New York, 1969.
8. Jullien, R., J. Phys. I (France) 2, 759 (1992). 13. Farias, T. L., Koylu, U. O., and Carvalho, M. G., Appl. Optics 35, 6560
9. Yanwei, Z., and Meriani, S., J. Appl. Cryst. 27, 782 (1994). (1996).
10. Bohren, C. F., and Huffman, D. R., “Absorption and Scattering of Light by 14. Heimenz, P. C., and Rajagopalan, R., “Principles of Colloid and Surface
Small Particles.” Wiley, New York, 1983. Chemistry,” 3rd ed., p. 218. Marcell Dekker, New York, 1997.

You might also like