You are on page 1of 12

w a t e r r e s e a r c h 4 6 ( 2 0 1 2 ) 3 1 7 7 e3 1 8 8

Available online at www.sciencedirect.com

journal homepage: www.elsevier.com/locate/watres

Electrocoagulation versus chemical coagulation: Coagulation/


flocculation mechanisms and resulting floc characteristics

Tali Harif*, Moti Khai, Avner Adin


Soil & Water Sciences Department, Robert H. Smith Faculty of Agricultural, Food and Environmental Quality Sciences, The Hebrew University
of Jerusalem, POB 12, Rehovot 71600, Israel

article info abstract

Article history: Electrocoagulation (EC) and chemical coagulation (CC) are employed in water treatment for
Received 1 December 2011 particle removal. Although both are used for similar purposes, they differ in their dosing
Received in revised form method e in EC the coagulant is added by electrolytic oxidation of an appropriate anode
29 February 2012 material, while in CC dissolution of a chemical coagulant is used. These different methods
Accepted 19 March 2012 in fact induce different chemical environments, which should impact coagulation/floccu-
Available online 29 March 2012 lation mechanisms and subsequent floc formation. Hence, the process implications when
choosing which to apply should be significant. This study elucidates differences in coag-
Keywords: ulation/flocculation mechanisms in EC versus CC and their subsequent effect on floc
Electrocoagulation growth kinetics and structural evolution. A buffered kaolin suspension served as a repre-
Aluminum sentative solution that underwent EC and CC by applying aluminum via additive dosing
Floc regime in batch mode. In EC an aluminum anode generated the active species while in CC,
Growth commercial alum was used. Aluminum equivalent doses were applied, at initial pH values
Zeta potential of 5, 6.5 and 8, while samples were taken over pre-determined time intervals, and analyzed
Scattering exponent for pH, particle size distribution, z potential, and structural properties. EC generated fragile
flocs, compared to CC, over a wider pH range, at a substantially higher growth rate, that
were prone to restructuring and compaction. The results suggest that the flocculation
mechanism governing EC in sweep floc conditions is of Diffusion Limited Cluster Aggre-
gation (DCLA) nature, versus a Reaction Limited Cluster Aggregation (RLCA) type in CC. The
implications of these differences are discussed.
ª 2012 Elsevier Ltd. All rights reserved.

1. Introduction by the reduction of repulsive forces between particles or by


the enmeshment in precipitates (Hogg, 2005). For insoluble
1.1. Coagulationeflocculation and factors determining particles, such as many minerals (e.g. Kaolin), inter-particle
floc evolution repulsion is usually due to electrical double layer interaction.
The addition of soluble ionic species will affect the surface
CoagulationeFlocculation in general is a two phase process potential (electrical potential difference between the particle
aimed at removing stable particles by forming larger aggre- surface and the bulk solution) of colloidal particles by either
gates that can be separated from the aqueous phase by adsorption to the particle surface or by double layer
a subsequent separation step. The preliminary phase is the compression. Ionic species that are specifically adsorbed at
coagulation phase in which destabilization is induced, either the surface can include multivalent cations and anions, ionic

* Corresponding author.
E-mail addresses: talih@atlantium.com (T. Harif), moty@ely.org.il (M. Khai), adin@vms.huji.ac.il (A. Adin).
0043-1354/$ e see front matter ª 2012 Elsevier Ltd. All rights reserved.
doi:10.1016/j.watres.2012.03.034
3178 w a t e r r e s e a r c h 4 6 ( 2 0 1 2 ) 3 1 7 7 e3 1 8 8

surfactants and, particularly in aqueous systems, hydrogen treatment method prior to membrane filtration, both for
and hydroxyl ions (Hogg, 2005). The type of coagulation achieving flux enhancement (Harif et al., 2006; Ben Sasson and
mechanism induced e be it surface adsorption or precipitate Adin, 2010), and for optimizing virus removal (Zhu et al., 2005).
formation (“sweep “floc”) e when using metal salts (such as Despite the research interest in EC, the vast majority of
aluminum or iron) is governed by the solubility boundary of studies have been conducted at an applied level, rather than at
the coagulant and therefore is highly dependent on pH, and a mechanistic one. Sporadic studies have explored some
coagulant dose (Amirtharajah and O’Melia, 1990). Flocculation mechanistic aspects of the process, distinguishing between
is the second stage of the combined process and consists of various coagulant species (Ben Sasson et al., 2009;
the aggregation of coagulated particles and/or precipitate Lakshmanan et al., 2009), and ascertaining how the unique
precursors in flocs. Successful inter-particle aggregation will chemical environment impacts the structural properties of
depend on the destabilization degree attained in the coagu- flocs (Harif and Adin, 2007, 2011). Comparative studies with
lation phase (the extent of a repulsive force barrier existing chemical flocculation (CC) are scarce, and limited to removal
between the particles), and on the collision rate between the efficiency comparative assessments (Holt et al., 2002; Bagga
particles. The latter is a function of hydrodynamics dictated et al., 2008; Cañizares et al., 2009). Hence, a more funda-
by system geometry and therefore can greatly differ between mental approach may be required to be able to postulate the
various systems (Bouyer et al., 2004). When referring to floc relevance of the EC process in water treatment processes
growth, several phases can be identified (Tambo, 1991). After versus the CC approach.
a characteristic time, a steady state is reached between In EC the active coagulant species are generated in situ by
aggregation and fragmentation, marked by a floc size distri- electrolytic oxidation of an appropriate anode material, thus
bution that is constant with time and is unique to the pre- differing from CC in which chemical coagulants such as metal
vailing shear conditions of each system (Spicer and Pratsinis, salts or polymers and polyelectrolytes are used. The half-cell
1996; Bouyer et al., 2004; Jarvis et al., 2005a). In general, the electrochemical reactions occurring in the EC cell (using an
overall floc growth rate (Rfloc) can be defined as (Jarvis et al., aluminum anode) and their corresponding standard electrode
2005a): potentials (written by convention as reductions) are:
Cathode:
Rfloc ¼ aRcol  Rbr (1)

Rcol ¼ rate of particle collision (frequency of collisions) 2H2 O þ 2e / H2ðgÞ þ 2OH E+ ¼ 0:83 V (2)
ðaqÞ
Rbr ¼ rate of aggregate breakage
a ¼ collision efficiency factor (fraction of collisions resulting in Anode:
attachment)
O2ðgÞ þ 4Hþ 
ðaqÞ þ 4e /2H2 O E+ ¼ þ1:229 V (3)

Floc growth will also be affected by the specific aggregate


AlðaqÞ þ 3e /AlðsÞ E+ ¼ 1:662 V

structure because a more open structure is considered to have (4)
a larger collision profile (Kusters et al., 1997). More porous flocs
are fragmented by fluid shear stresses more rapidly than When considering the half-reaction standard potentials,
compact mass equivalent flocs (Potanin, 1991; Flesch et al., anodic Al3þ dissolution yields a significantly lower Gibbs free
1999). Restructuring can also occur, by re-aggregation of energy compared to anodic water reduction, therefore Al3þ
fragments or by shear interactions that rearrange the struc- dissolution will ultimately dominate. Assuming that a minor
ture (Oles, 1992; Jarvis et al., 2005b). Two types of aggregation portion of the current will go toward anodic water reduction
have been identified: diffusion limited cluster aggregation the outcome is that hydroxyl ions (OH) will be formed at the
(DLCA) and reaction limited cluster aggregation (RLCA) (Tang cathode in excess in relation to hydrogen ions (Hþ) at the
et al., 2000; Bushell et al., 2002). The former occurs when the anode. Although Al(OH)3 precipitation can remove some OH
repulsive barrier between particles is low e forming looser, this is not suffice to counteract the overall OH production.
tenuous structures; the latter occurs when the repulsive The rate of OH formation surpasses the rate of Al(OH)3
barrier between particles is high e producing more compact precipitation, even in optimal precipitation conditions,
and stronger structures (Tang et al., 2000). particularly when considering that 1 mole of produced Al3þ
will divide into an array of hydrolyzed species, some of
1.2. Electrocoagulation in water treatment which are soluble. Hence, the overall reactions result in
excess OH in solution, manifested in a transient pH rise
Electrocoagulation (EC) over the past decade has gained over time. This phenomenon has previously been reported
recognition as an effective process for various water treat- (Chen et al., 2000; Harif and Adin, 2007, 2011; Bagga et al.,
ment applications. Research into EC, although not extensive, 2008; Lakshmanan et al., 2009; Cañizares et al., 2009) and
has examined colloidal and organic matter removal (Vik et al., differs greatly from CC, in which the pH decreases. Coagu-
1984; Matteson et al., 1995; Holt et al., 2002; Larue et al., 2003) lation mechanisms depend primarily on pH and coagulant
and explored a wide range of applications in urban and dosage, which govern speciation of the active mononuclear
industrial wastewater treatment (Ogutveren and Koparal, species (Amirtharajah and O’Melia, 1990; Sposito, 1996;
1992; Alexandrova et al., 1994; Pouet and Grasmick, 1995; Letterman et al., 1999; Duan and Gregory, 2003), thus such
Belongia et al., 1999; Mollah et al., 2001, 2004; Adin and Vescan, a difference is of pivotal importance when comparing both
2002). Additionally, EC has been found to be an effective pre- processes. Disparities may also exist in the hydrolysis of the
w a t e r r e s e a r c h 4 6 ( 2 0 1 2 ) 3 1 7 7 e3 1 8 8 3179

aluminum species. In EC, as opposed to CC, negative aggregates. Calculating fractal structures derived from fractal
counter-ions are not added. Negative counter-ions can mathematics (Mandelbrot, 1982) is extensively discussed
influence the aluminum hydrolyzed species, because they elsewhere (Bushell et al., 2002; Jarvis et al., 2005b).
can replace the hydroxyl (Duan and Gregory, 2003) and their Using the fractal dimension as a measure of structural
absence in EC may affect the coagulation mechanisms. The measurements is applicable if the RayleigheGanseDebye
objective of this paper is to elucidate fundamental chemical (RGD) approximation is met, for non-absorbing particles in the
differences existing between EC and CC that would impact limit that both:
coagulation mechanisms, flocculation behavior and subse-
quent floc characteristics. jm  1j  1 (6a)

1.2.1. Electrocoagulation technology and coagulant types ð4pn=lÞL jm  1j  1 (6b)


The most basic EC reactor may be made up of an electrolytic where m is the relative refractive index of the scatterers
cell containing one anode and one cathode. The anode metals (primary particles) and L is the length of the scattering body
most commonly used are aluminum or iron because when (the diameter of the primary particles).
electrochemically oxidized they produce the most commonly For amorphous Al(OH)3 nucleation clusters (hereafter
used ionic coagulants, Al3þ and Fe3þ (or Fe2þ) respectively. The termed “Al(OH)3 precursors”), a refractive index of 1.59 can be
dissolution of coagulant into solution is governed by Faraday’s used as an approximation, based on the refractive index range
Law: of other Al(OH)3 solid forms (Li et al., 2005; Harif and Adin,
2011), thereby fulfilling the requirements in equations (6a)
ItM
w¼ (5) and (6b).
ZF
The fractal dimension of an aggregate, assuming RGD
w ¼ metal dissolving (gr M/cm2) approximation, is acquired from the slope of the logarithmic
I ¼ current intensity (A) plot of the angular scattering intensity (I(q)) versus the
t ¼ time (s) momentum transfer (q). However, due to variations in aggre-
M ¼ molecular weight of metal (gr/mol) gate structures and sizes comprising the entire scattering
Z ¼ number of electrons involved in the oxidation/reduction volume, the absolute slope of the logarithmic plot of I(q)
reaction versus q will be referred to as the scattering exponent (SE), as it
F ¼ Faraday’s constant (96,485 C) may not necessarily represent the real mass fractal dimension
of all the aggregates. The SE, however, does still portray
Aluminum hydrolysis and mononuclear species formation structural properties, and higher SE values will indicate more
following aluminum dosing has been discussed extensively in compact aggregates e and vice versa (Guan et al., 1998; Waite,
Amirtharajah and O’Melia (1990) and Duan and Gregory (2003). 1999).
The solubility boundary of aluminum (0.03 mg/l Al3þ, at pH The system presented in this study represents a complex
6.3) denotes the thermodynamic equilibrium that exists water system, thus the light scattering results presented for
between the dominant aluminum species and amorphous a mixture of Al(OH)3 precursors and kaolin should be inter-
aluminum hydroxide, Al(OH)3, (the assumed solid form rele- preted with caution. However, for relatively narrow size
vant in coagulation processes) at a given pH. Although distributions, polydispersity effects are expected to be insig-
dimeric, trimeric and polynuclear hydrolysis products of Al3þ nificant (Lawler, 1997; Bushell and Amal, 1998).
can form, these can often be ignored, especially in dilute
solutions, and may not affect the overall speciation (Duan and
Gregory, 2003). As such, this paper will assume that mono- 2. Materials and methods
nuclear hydrolyzed species adequately predict Al(OH)3
precipitation. 2.1. Colloidal suspension

1.2 g of kaolin (AlSi2O5(OH)4, Aldrich Chemical Company Inc.


1.3. Floc structural interpretation
USA) was suspended in 20 L of distilled water (60 mg/l final
concentration), dispersed and homogenized using an Ultra-
Characterization of structures can be done using various
turax 2000 (Ganke and Kunkel, GMBH). 1.66 g of NaHCO3 (BDH
techniques which are reviewed in Jarvis et al. (2005b). Static
Laboratory Suppliers, UK) was added (final concentration
light scattering will be used in this research, and is considered
83 mg/l), the pH was corrected to 5, 6.5 and 8 with NaOH or
advantageous due to the non-interfering nature of the tech-
H2SO4, and conductivity was increased to 1 mS/cm with
nique (Thill et al., 2000).
NaNO3 (Riedel-Dehaen, GMBH).

1.3.1. Light scattering techniques


Particle scattering patterns and intensities can be translated 2.2. Coagulation cells
into a measure of the particle size, according to available
approximations that were developed for particles of different 2.2.1. Electrocoagulation cell
sizes and optical properties (Sorensen, 1997). The scattering An EC batch cell was designed. It consisted of a plexyglass cap,
patterns are related to the scattering angle in a way which can which could be fitted onto a 1 L chemical glass and to which
be used to obtain information also on the structure of the the electrodes were attached. The inner electrode, the
3180 w a t e r r e s e a r c h 4 6 ( 2 0 1 2 ) 3 1 7 7 e3 1 8 8

cathode, was made from stainless steel and concentric in the electrostatic forces of repulsion between charged particles
form (H ¼ 10 cm, D ¼ 2.5 cm). It was fitted onto the arm of the which changes with the addition of a coagulant.
magnetic stirrer, and stationary throughout all the experi-
ments. The outer electrode, the anode, was made from 2.6. Experimental procedure
aluminum, designed, also concentric in form with dimen-
sions: H ¼ 10 cm, D ¼ 9.5 cm. The electrodes were connected to 2.6.1. Electrocoagulation
a DC external power source. The EC apparatus was fixed onto a 1 L chemical glass con-
taining 800 ml of kaolin suspension, the electrodes submerged
2.2.2. Chemical coagulation cell in the suspension, and magnetic stirrer set at a constant
To attain comparative chemical conditions between the gentle speed of 145 rpm. The stirring speed was selected to
processes, for each time increment, alum was added to the induce floc growth over the defined experimental time range,
suspension for identical time intervals as current operation in based on preliminary experiments. EC was conducted in gal-
the EC cell. A 1 L chemical glass was used, fitted with the vanostatic mode: the current was set and the potential found
identical plexyglass cap and magnetic stirrer used in the EC its own value dependent on the system’s overall resistance.
cell, and a stock solution of commercial alum (Al2(SO4)318H2O, This ensured coagulant production at a pre-determined rate,
SigmaeAldrich, Israel) was fed using a syringe pump (74900 defined by Faraday’s law.
series, ColeeParmer, USA). Preliminary experiments were conducted to ensure that
the aluminum dissolution could be calculated by Faraday’s
2.3. Floc size and structure analysis Law. Fig. 1 shows a comparison between the theoretical and
actual aluminum dissolution, as a function of time, in an EC
Size distributions and structural information of kaolin- cell using an aluminum anode and stainless steel cathode.
Al(OH)3 flocs were determined as a function of time using Aluminum dissolution was measured by weighing the anode
a Mastersizer Microplus (Malvern Instruments, UK), which at each time interval.
ascertains size by analysis of forward scattered light. The size The graph shows a nearly ideal correlation between the
distribution data given by the instrument covers the size theoretical calculation, using Faraday’s Law, and the actual
range of 50 nme500 mm. Size distribution information was values obtained. These values cover much higher aluminum
obtained using supplied software which uses Mie theory to doses and longer time spans than applied in this research,
develop a scattering pattern that matches the scattering thus we can conclude that the set-up can generate a target a-
pattern of the sample being measured. Information on mean luminum dose. A stainless steel cathode was chosen to
distribution size is presented in this paper as the volume minimize effects of cathodic dissolution that would lead to
mean diameter: a positive deviation from the theoretical Faradaic values. The
X X topic of cathodic dissolution in EC cells and ways to overcome
DðV; 0:5Þ ¼ ðVi di Þ= Vi (7) this phenomenon are discussed in Ivanshivili et al. (1987) and
i i
Picard et al. (2000).
Where Vi is the relative volume in size class i with mean class The mixing regime applied was chosen to enable discrete
diameter di. monitoring of floc evolution during coagulation and initial
Information on floc structure was obtained by measuring growth stages and provided the platform for analogous
the intensity of light (I ) at all detectors, for three consecutive fundamental evaluation. At various time intervals (0, 3, 6 and
runs, and plotting log I versus log q. Information regarding the 10 min) samples were taken for analyses (pH, size distribution,
angles of the detectors and intensity correction data, based on
the geometric configuration of the detectors, was supplied by
Malvern Instruments.

2.4. Image analysis

Image analysis was used in conjunction with scattering


measurements as a complementary analysis to ascertain
qualitatively floc properties over time. The flocculated
suspension was gently poured into a petri dish and the flocs
were photographed using a digital camera (model DP11,
Olympus, Japan) which was mounted on a Stereoscope (model
SZX12, Olympus, Japan). All photographs were taken at 3.0
mega pixel resolution, and magnification 50.

2.5. z potential

The z potential of the suspension was measured using Fig. 1 e A comparison between theoretical and actual
a Zetamaster S (Malvern Instruments, UK). Each result was an aluminum anodic dissolution in the EC cell, using an
average of three readings. Its value determines the extent of aluminum anode and stainless steel cathode.
w a t e r r e s e a r c h 4 6 ( 2 0 1 2 ) 3 1 7 7 e3 1 8 8 3181

z potential and image analysis). The complete dose of 300 mm. In CC, at pH 8, floc formation was observed only with
aluminum was achieved at 10 min, for each applied current, 160 mg/l alum. The growth histogram indicates a very short
after which additional aluminum was not introduced into induction period (up to 3 min), after which slower growth is
solution. At optimal sweep floc conditions (initial pH 6.5) observed, with final modal diameters reaching 250 mm. In EC,
samples underwent continuous mixing and measurements at pH 8, floc formation did not occur, for all currents applied.
were taken again at 15 min. For all size distribution measure- The histograms attained in EC at pH 5 show a long induc-
ments, the Mastersizer Microplus was operated at the lowest tion phase, and a paced type of growth. Interestingly, despite
possible pump speed of 400 rpm, so to minimize disruption of this, the final modal diameters attained were larger than those
floc structure. Qualitative analysis using image analysis obtained in CC at pH 8 with 160 mg/l alum; 300 mm versus
techniques corroborated relative floc sizes attained. The 250 mm, for EC and CC respectively.
currents used for the experiments were 0.042 A, 0.11 A and The results suggest a low collision frequency and efficiency
0.22 A, yielding aluminum doses of 2.4 mg/l, 6.5 mg/l and for EC at pH 8 compared to CC, however at pH 5, this is
13 mg/l respectively. These doses are equivalent to aluminum reversed. The collision frequency is a function of precursor
content in 30 mg/l, 80 mg/l and 160 mg/l commercial alum particle concentration (kaolin primary particle concentration,
(Al2(SO4)318H2O) used in the CC experiments (8.1% of the total and precipitation of Al(OH)3 precursors formed by addition of
molecule). aluminum coagulant into solution), aggregate geometry/
collision profile, and the latter a function of repulsive elec-
2.6.2. Chemical Coagulation trostatic barriers (Kusters et al., 1997; Chakraborti et al., 2003).
A 1 L chemical glass containing 800 ml of kaolin suspension The pH and z potential measured over each time increment in
was fitted with the plexyglass cap and magnetic stirrer set at these experiments (Fig. 2) can shed light on the different
145 rpm (as in all EC experiments). A stock solution of growth behavior observed between both types of coagulation.
10  103 mg/l commercial alum (Al2(SO4)318H2O) was prepared The first most prominent difference observed between EC
and fed at a preset dosing rate of 0.24 ml/min, 0.64 ml/min and and CC pertains to the pH change over time. In general, in EC
1.28 ml/min, to achieve final doses of 30 mg/l, 80 mg/l, and the pH rises, whereas in CC the pH decreases. This phenom-
160 mg/l respectively. The alum was introduced into solution enon and its significance have been discussed previously.
over the identical time span as current operation in the EC cell At pH 8, in EC, the pH did not change for the duration of the
(10 min). In optimal sweep floc conditions (initial pH 6.5), the experiments, for both currents applied. The z potential of the
suspension underwent continuous mixing for an additional solution also was stable, and maintained a value of approxi-
5 min. Samples were taken for analyses (pH, size distribution, mately 40 mV. Both z potential and pH stabilization at pH 8
z potential and image analysis) at 0, 3, 6, 10 and in optimal suggest the removal of excess OH from solution. Removal
sweep floc conditions also at 15 min. could be attributed to the buffering capacity of the solution,
The mixing conditions applied both in the EC cell and CC which was maintained due to bicarbonate predominating at
cell were aimed at inducing floc formation. A range of average pH 8 under open atmosphere, and also to formation of
velocity gradients have been used in various studies exam- neutrally charged species, Al(OH)3. The formation of Al(OH)3 in
ining the impact of shear stress on floc growth, and empirical these conditions would be primarily dictated by Le Chatelier’s
relationships developed (Jarvis et al., 2006). However, devia- principle, due to excess OH cathodic formation, which drives
tions between the average velocity gradient and resulting floc the solubility equilibrium toward precipitation. The overall
sizes are apparent, while it is known that a specific floc growth removal of OH from solution, maintains the pH at a stable
curve is unique to each system (Spicer et al., 1998; Bouyer value, inhibiting a pH rise and subsequent formation of
et al., 2004; Jarvis et al., 2005a). Furthermore, studies have negatively charged species (Al(OH)4), which otherwise would
suggested that local velocity gradients have a more significant have lowered the z potential to a more negative value. The
impact then average velocity gradients leading investigators coupled effect obstructs the growth of Al(OH)3 precursors into
to adapt the use of rpm instead of a calculated average shear larger flocs, due to the high repulsive electrostatic barrier
velocity (Jarvis et al., 2006). In light of this, the results obtained which exists in solution (reflected by the very negative z
from using a specific mixing regime, must be limited to the potential). On the other hand, when applying CC at pH 8, the
set-up. In this study, the results are comparative and relevant pH is lowered, and reaches 5.8 for a dose of 160 mg/l alum. Floc
to a mixing regime that can be considered representative and growth was not evident at pH 8 for lower doses of alum. The pH
falls within the wide range of velocity gradients applied for value reached was not low enough to induce adequate positive
attaining floc growth. mononuclear species (AlOH2þ, Al(OH)þ 2 ) formation, which
would have lowered the electrostatic repulsive barrier, depic-
ted by a shift in the z potential toward a more positive value.
3. Results and discussion Conditions in which both the electrostatic repulsive barrier is
adequately lowered (reflected by a more positive z potential
3.1. Size distributions and floc formation pH 5 and 8 value of 10 mV), and Al(OH)3 precursor mass production is
suffice, occurs when applying a high alum dose. The induction
Figs. S1 and S2 in supporting information show representative period is short, indicating high collision efficiencies at the early
growth histograms obtained with CC and EC respectively. In stages, however the growth kinetics show that as the floccu-
CC, at pH 5, no substantial floc growth was observed. On the lation progresses, the rate of growth slows, suggesting that
other hand, in EC, at pH 5, all currents yielded flocs of a steady state between aggregation and fragmentation is being
considerable size, with modal diameters reaching above reached (Oles, 1992; Spicer et al., 1998; Jarvis et al., 2005a).
3182 w a t e r r e s e a r c h 4 6 ( 2 0 1 2 ) 3 1 7 7 e3 1 8 8

Fig. 2 e A. pH changes at initial pH 5, B. pH changes at initial pH 8, C. z potential changes at initial pH 5, D. z potential


changes at initial pH 8. Each z potential result is an average of three consecutive readings.

At pH 5 positive aluminum mononuclear species (AlOH2þ, precursor formation is nearly optimal as the minimum solu-
Al(OH)þ 2 ) govern the solution. In EC, the pH rises substantially bility of Al(OH)3 occurs at pH 6.3 (Amirtharajah and O’Melia,
and conditions become favorable for Al(OH)3 precipitation 1990). As such, these conditions were chosen to serve as
(transition into sweep-floc regime), subsequently causing an a basis for structural comparisons between flocs obtained in
acceleration in floc growth at later stages. The reason for pH each process, and an additional 5 min mixing was used to
rise is excess cathodic production of OH, and in addition the attain longer growth periods. Representative growth histo-
lowered buffering capacity of the solution at pH 5, due to grams are shown in Fig. 3.
transition toward CO2 gas formation (originating from bicar- Floc size evolution over time show for all applied currents
bonate), which diffuses out of solution under atmospheric and lower alum doses growth up to modal diameters of
pressure. The size distribution spreads indicate that smaller approximately 380 mm, with the size distribution spread
particles are being formed simultaneously, and are able to reaching the upper detection limit of the instrument. Previous
aggregate more efficiently into larger flocs at higher currents. studies examining alum induced flocculation of colloidal
For both currents at initial pH 5, the z potential exhibits charge suspensions have documented modal diameters from 200 mm
reversal, with similar values reaching þ18 mV at the termi- (Bouyer et al., 2004) to 500 mm (Spicer et al., 1998), following
nation of the process (10 min). The rather large disparities in z 15 min of slow mixing. The floc sizes attained in this study are
potential attained at initial pH 5 between EC and CC are most well within that range, hence we can assume that the additive
likely due to the soluble negative ions introduced with the dosing regime applied in CC was able to generate floc sizes
chemical coagulant that mask the positive ion affect thus compatible to conventionally practiced methods. For a high
limiting the z potential from reaching very positive values. alum dose the modal diameter obtained at 15 min is lower,
Despite the higher repulsive barrier, EC was able to generate and reaches 200 mm. For 0.042 A the induction phase is longer
flocs, suggesting that floc growth at this pH is derived from than that observed for 0.22 A, but at 6 min accelerates, yielding
increased Al(OH)3 precipitate mass formation at higher modal diameters similar to those attained with the high
currents leading to higher collision frequencies. In CC, despite current (above 300 mm). For 0.22 A, acceleration in growth at
attaining a lower repulsive barrier, reflected by a near iso- 6 min is also observed, resulting in near maximum modal
electric point z potential, precursor formation is limited, as diameters by 6 min. The growth inhibition evident at 10 min
the pH values inhibit Al(OH)3 precipitation. and above, is due to the enhanced breakage rate, dictated by
shear forces which restrict evolution into a larger floc.
3.1.1. pH 6.5 EC produces narrower size distributions at 10 and 15 min
At initial pH 6.5, both CC and EC yielded sufficient floc growth compared to CC, while both processes exhibit wide size
for all alum doses and all currents applied. At this pH, Al(OH)3 distributions up to 6 min. The wide size distributions indicate
w a t e r r e s e a r c h 4 6 ( 2 0 1 2 ) 3 1 7 7 e3 1 8 8 3183

Fig. 3 e Growth histograms obtained with low and high alum doses and applied currents at initial pH 6.5. Coagulant
addition occurred over the first 10 min, following an additional 5 min mixing period, in aim of attaining longer growth
patterns. The size distribution at 0 min is that of the initial kaolin suspension.

8
that while growth into larger floc is occurring, smaller parti- A 0.042A
cles are simultaneously being formed. For 0.22 A growth 7 0.11A
acceleration occurs at 3 min, while for 0.042 A at 6 min, indi-
cating a linear trend between current applied and floc 0.22A
6
pH

formation. CC, on the other hand, attains floc sizes similar to


those reached in EC at a slower rate (at 10 min and above as 30mg/l
5
opposed to 3 and 6 min in EC). These differences can be
80mg/l
explained by Al(OH)3 precursor formation, pH changes and z
potential changes over time. Fig. 4 shows the differences in z 4 160mg/l
potential and pH values for each time increment, between EC 0 5 10 15
Time (min)
and CC at initial pH 6.5.
The differences in pH and z potential values over time in Time (min)
both processes underline the differences in the physico- 0 5 10 15
chemical environments both induce. For alum dosing of 30
80 mg/l and 160 mg/l the pH rapidly decreases at 6 min dosing,
B 0.042A
20
0.11A
ζ Potential (mV)

reaching pH values below 5, by the end of the process. Dosing


10
of 30 mg/l alum, however, lowers the pH marginally, stabi-
lizing at a value of 6 during the final stages of the process. In 0 0.22A
these conditions a larger number of Al(OH)3 precursors can -10 30mg/l
form, as opposed to the higher alum doses, because all growth
-20 80mg/l
stages occur within nearly optimal sweep-floc regime; hence
the increased growth and larger final sizes. Despite a stronger -30
160mg/l
repulsive barrier (reflected by a more positive z potential)
-40
exhibited in CC versus EC, growth is appreciable, and for
30 mg/l reaches a maximum modal diameter of 380 mm. For Fig. 4 e EC and CC at initial pH 6.5 A: Changes in pH values
the higher alum doses transition out of sweep-floc regime over time for all currents and alum doses applied. B:
toward positive mononuclear speciation is pronounced, Changes in z potential values over time for all currents and
limiting mass precursor formation in the overall process, alum doses applied. Each z potential result is an average of
which in turn, limits floc growth. three consecutive readings.
3184 w a t e r r e s e a r c h 4 6 ( 2 0 1 2 ) 3 1 7 7 e3 1 8 8

In EC, the z potential values following application of current correlation between growth and dose was not observed, and
approach the iso-electric point, indicating a diminished elec- for sweep floc conditions, the lowest dose yielded both faster
trostatic repulsive barrier. As such, the limiting factor for floc growth rates and larger final sizes. For 30 mg/l alum the final
growth in EC at initial pH 6.5 is apparently a decrease in colli- volume mean diameters are similar to those attained for
sion frequencies, due to transition out of sweep-floc regime. 0.22 A, but the path is somewhat different. For 0.22 A,
The pH rise, stemming from excess OH cathodic production a distinct sigmoidal profile is observed, showing pronounced
and lowered buffering capacity (due to CO2 gas diffusion out of fragmentation at 6 min, indicating transition into steady state
solution under atmospheric pressure following the initial pH flocculation (Spicer et al., 1998; Bouyer et al., 2004; Jarvis et al.,
correction) indicates that the precipitation potential of Al(OH)3 2005a). For 30 mg/l the growth profile is also sigmoidal, with
decreases over time, resulting in less OH removal. The a mild deceleration observed around 12 min, indicating frag-
enhanced production of Al(OH)3 precursors at the initial stages mentation only at later stages. The growth curves of the flocs
of the process, most predominant at 0.22 A (exhibiting a very obtained from applying smaller currents show also sigmoidal
short induction period), enables adequate precursor mass behavior e fragmentation occurring at 6 min and 10 min for
formation that lay the foundation for appreciable floc evolu- 0.11 A and 0.042 A respectively. Alum dosing at higher doses
tion, despite the transition into unfavorable precipitation produced growth curves devoid of sigmoidal behavior, yet
conditions, as the process progresses. suited to polynomial growth. In CC of a kaolin clay suspension
using alum, the rate determining step is dictated by inter-
particle collision (Matsui et al., 1998). It seems that in EC,
3.2. Floc growth profiles at pH 6.5
this step is shortened, due to higher Al(OH)3 precursor mass
production leading to a higher collision frequency. This,
Gaining a deeper understanding of specific growth kinetic
coupled with a lower repulsive barrier, induces a rapid growth
patterns between both processes in sweep-floc range requires
rate. In CC, the floc growth is indeed slower, suggesting
plotting the volume mean diameter against time. Fig. 5 shows
smaller collision profiles. The flocs formed in EC apparently
the evolution of volume mean diameter for all currents and
possess a weaker strength and are more susceptible to frag-
alum doses applied at initial pH 6.5, and interestingly unveils
mentation. For sigmoidal growth curves, transition into
distinct differences in growth kinetics between EC and CC.
steady state occurred in EC at approximately 220 mm, 180 mm
Significant variations between both processes are evident.
and 200 mm for 0.22 A, 0.11 A and 0.042 A respectively. This
The particular growth patterns representative of each process
occurred in CC (30 mg/l) at approximately 260 mm. Floc
are a product of solution chemistry that gives rise to different
strength is indicated by the size of a floc attained at a given
floc sizes, structures, strengths and subsequent collision
shear rate (Jarvis et al., 2005a). The general understanding is
profiles, the importance of which has been presented else-
that strength increases with increasing size, a relationship
where (Potanin, 1991; Oles, 1992; Spicer and Pratsinis, 1996;
found to be particularly appropriate for alumino-clay and
Kusters et al., 1997; Flesch et al., 1999; Waite, 1999; Chakraborti
alumino-humic flocs (Bache and Gregory, 2010). The strength
et al., 2003). EC shows distinctly faster growth rates, compared
referred to is for the given shear condition under which the
to CC, for all applied currents and alum dosing concentrations.
flocs were formed. In addition to floc size, floc structural
In EC, a higher dose produces both a faster growth rate and
evolution can be indicative of floc strength, as the floc bonds
larger final floc sizes. In CC, on the other hand, a linear
(numbers and strength) will dictate a floc’s propensity to
undergo restructuring. Hence, complementary structural
analysis was performed to attain a comprehensive picture of
specific flocculation mechanisms and floc structural behavior,
which are apparently individual to each process.

3.3. Floc structural evolution at pH 6.5

Typical scattering plots (I(arb) versus q(nm1)) were used to


obtain the structural data. Interpretation of these plot types is
explained elsewhere (Thill et al., 2000; Bushell et al., 2002). For
all scattering graphs, one decade of linearity was observed for
1.4*104 < q < 1.6*103 (Figs. S3 and S4 in supporting infor-
mation). This region was used to calculate the scattering
exponent (SE) from the graph slopes. Table 1 summarizes the
SE values obtained.
The flocs formed with all currents applied in EC, under-
Fig. 5 e Evolution of volume mean diameter for all currents went compaction resulting in a final SE, at 15 min, which is
and alum doses at pH 6.5. The fitted curves show for all higher than that obtained during the initial 3 min. In some
currents and lowest alum dose a sigmoidal growth pattern, cases, fluctuations in the structural evolution are apparent,
indicating a fragmentation stage, and for higher alum suggesting the flocs are of fragile nature, and are prone to
doses, a second order polynomial pattern, indicating no fragmentation and restructuring, and ultimately compaction.
fragmentation. All coefficients of determination (R2) are in In CC, on the other hand, for 30 mg/l alum, at which appre-
the range of 0.98e0.99. ciable growth is observed, the structural evolution shows that
w a t e r r e s e a r c h 4 6 ( 2 0 1 2 ) 3 1 7 7 e3 1 8 8 3185

Table 1 e Scattering exponent values for time increments over 3e15 min, obtained at pH 6.5 for all currents and alum doses
applied. The higher values indicate a more compact structure and the lower values a more porous structure.
Time (min) Scattering exponent

30 mg/l Alum 0.042 A 80 mg/l Alum 0.11 A 160 mg/l Alum 0.22 A

3 1.99  0.03 1.83  0.04 1.87  0.02 1.78  0.02 1.81  0.03 1.77  0.02
6 1.91  0.02 1.93  0.01 1.89  0.01 1.95  0.01 1.84  0.01 1.78  0.01
10 1.83  0.01 1.89  0.02 1.82  0.01 1.88  0.02 1.84  0.01 1.81  0.02
15 1.78  0.02 1.93  0.01 1.84  0.01 1.93  0.01 1.86  0.01 1.86  0.01

at initial stages of flocculation, a more compact structure readily upon contact and form tenuous structures (Tang et al.,
exists that becomes more porous, as the flocculation prog- 2000). These structures have higher collision profiles, yet
resses. Both favorable Al(OH)3 precipitation conditions and are prone to fragmentation and restructuring. As such, in EC
surface forces enable growth into a larger floc, that exhibit the repulsive barrier is indeed negligible thus enabling accel-
a stability not observed for equivalent sized flocs obtained in erated growth. However, the large tenuous structure, and
EC. The SE values obtained for the higher alum doses change subsequent large collision profile at the early stages, seems
insignificantly throughout the process, maintaining generally fragile and more susceptible to the applied shear force, hence
a stable value, within a margin of error. These doses also floc restructuring occurs, leading to a more compact form by
generated slow growth profiles, and smallest floc sizes, due to the end of the process. The gaseous products (hydrogen and
smaller collision profiles, because of growth limitation oxygen) formed in EC process could also facilitate the adhe-
dictated by a high repulsive barrier and transition into inad- sion of precursor particles together. Bubble nucleation rate
equate Al(OH)3 precipitation conditions. Comparison of SE increases with increasing currents (Shahjahan Kaisar Alam
values arising from conditions in which floc growth generated Sarkar et al., 2010), and indeed could impact the linear corre-
comparable floc sizes (30 mg/l alum versus 0.22 A), indicates lation existing between current applied and floc growth in EC.
different aggregation regimes, dictated by different collision RLCA occurs when a substantial repulsive force remains
mechanisms. In EC, it seems that the flocs form by a DLCA between particles, so the “sticking” probability diminishes
type of reaction, whereas in CC the aggregation is more of an and particles rely on many collision frequencies to form
RLCA nature. DLCA occurs when there are negligible repulsive a stable floc. In this case, the floc formed will exhibit a tighter
forces between colloid particles, causing particles to “stick” and more compact structure. In CC the repulsive barrier is

Fig. 6 e A visual comparison of floc images formed at 6 and minutes with 30 mg/l alum dosing and 0.22 A. Magnification
503, a comparative scale of 500 mm is shown in the lower right image.
3186 w a t e r r e s e a r c h 4 6 ( 2 0 1 2 ) 3 1 7 7 e3 1 8 8

indeed higher, leading to a slower growth curve, yet evolution evolution patterns in EC, obtained in optimal sweep floc
of a more strong and compact floc, able to withstand shear conditions (initial pH 6.5) point to a DLCA mechanism versus
forces. The result of this was structural evolution into a final an RLCA type in CC. On a practical note, EC can diminish the
more porous structure than attained in EC. need for pH adjustment at low initial pH values, and could be
Structural evolution, as floc size evolution, is dependent specifically suited to processes in which shorter flocculation
not only on solution chemistry (dictated by the coagulant dose times are applied and initially a more porous (less dense)
and water characteristics) but also on specific flocculation aggregate is favored (i.e. dead-end/cake filtration, dewatering,
conditions (shear and time). Cycled shear flocculation and flotation). Smaller and more compact structures, less prone to
tapered shear flocculation have shown to yield distinct shear stress over longer flocculation times, like those gener-
structural differences (Spicer et al., 1998). The mixing condi- ated in CC, would most likely be suited to processes incorpo-
tions in this study are assumed to be indicative of alum rating higher shear environments. In line of this, particular
induced flocculation at a constant applied shear. Therefore attention should be made when choosing to use either EC or
the structural differences documented are relevant to the floc CC, depending on treatment process and overall objectives.
growth phase and transition into steady state.
Qualitative image analysis of the flocs corroborated the
scattering structural analysis. The flocs formed by alum
Acknowledgments
dosing are indeed more compact, well defined and denser
compared to the flocs formed by applied current, which
The work was partially supported by BMBF Germany and
appear “fluffy” and transparent. Fig. 6 shows visual differ-
Israeli Ministry of Science. We gratefully acknowledge Dr.
ences in floc structure at 6 and 10 min, using 30 mg/l alum
Rivka Amit and Mr. Yoav Nachmias from the Geological
versus 0.22 A.
Survey of Israel for assistance with the Malvern Mastersizer
Microplus, and additional instrumentation. Prof. Daniel
Mandler from the Hebrew University of Jerusalem is kindly
4. Concluding remarks
thanked for fruitful discussions.

Although EC and CC are considered similar coagulation


processes, targeted toward particle removal, in fact substan-
tial differences exist between them in terms of coagulation- Appendix A. Supporting information
flocculation mechanisms and floc evolution patterns. The
conclusions presented are based on a comparative evaluation Supplementary data related to this article can be found online
of solution chemistry and kaolin-Al(OH)3 floc evolution at doi:10.1016/j.watres.2012.03.034.
throughout growth stages. The authors acknowledge that the
choice of an additive dosing regime for CC is not common
references
practice, however, inducing time discrepancies between
dosing regimes would result in divergent coagulant concen-
trations, and subsequent coagulation-flocculation conditions
Amirtharajah, A., O’Melia, C.R., 1990. Coagulation processes:
in which disparities would not be comparable at the mecha-
destabilization, mixing and flocculation. In: Pontius, F.W. (Ed.),
nistic level. In effect, the applied CC dosing regime did Water Quality and Treatment, fourth ed. McGraw-Hill Inc.,
produce flocs comparable to other studies using the standard New-York.
method. In general, most studies exploring floc evolution Adin, A., Vescan, N., 2002. Electroflocculation for particle
using chemical coagulation have limited the findings to the destabilization and aggregation for municipal water and
unique experimental set-up because different systems have wastewater treatment. Proceedings of the American Chemical
Society 42 (2), 537e541.
shown to produce widely variable results with regard to floc
Alexandrova, L., Nedialkova, T., Nishkov, I., 1994.
formation rate and structural characteristics (Spicer et al., Electrofloatation of metal ions in waste water. International
1998; Bouyer et al., 2004; Jarvis et al., 2005a; Bache and Journal of Mineral Processing 41, 285e294.
Gregory, 2010). Notwithstanding, we postulate that the Bache, D.H., Gregory, R., 2010. Flocs and separation processes in
comparative supporting evidence provided can be projected to drinking water treatment: a review. Journal of Water Supply:
larger systems. The overall distinct solution chemistry diver- Research and Technology AQUA 59 (1), 16e30.
gences observed e a function of the particular coagulant Bagga, A., Chellam, S., Clifford, D.A., 2008. Evaluation of iron
chemical coagulation and electrocoagulation pretreatment for
addition method e are of a magnitude that heavily impact
surface water microfiltration. Journal of Membrane Science
coagulation mechanisms. As the coagulation mechanism 309, 82e93.
attained is the core driver of floc evolution in growth phase Belongia, B.M., Haworth, P.D., Baygents, J.C., Raghavan, S., 1999.
conditions, the implications of using EC or CC should be Treatment of alumina and silica chemical mechanical
significant. polishing waste by electrodecantation and electrocoagulation.
For equivalent coagulationeflocculation conditions, EC is Journal of the Electrochemical Society 146 (11), 4124e4130.
Ben Sasson, M., Adin, A., 2010. Fouling mechanisms and energy
able to produce flocs over a wider range of pH values relevant
appraisal in microfiltration pretreated by aluminum-based
to water treatment, apparently at a more rapid rate. Floc
electroflocculation. Journal of Membrane Science 352, 86e94.
structural evolution in EC suggests the formation of structures Ben Sasson, M., Calmano, W., Adin, A., 2009. Iron oxidation
that are relatively porous and fragile, prone to restructuring processes in an electroflocculation (electrocoagulation) cell.
and compaction. Both the floc formation rate and structural Journal of Hazardous Materials 171, 704e709.
w a t e r r e s e a r c h 4 6 ( 2 0 1 2 ) 3 1 7 7 e3 1 8 8 3187

Bouyer, D., Liné, A., Do-Quang, Z., 2004. Experimental analysis of Lawler, D.F., 1997. Particle size distributions in treatment
floc size distribution under different hydrodynamics in processes: theory and practice. Water Science and Technology
a mixing tank. AIChE Journal 50 (9), 2064e2081. 36 (4), 15e23.
Bushell, G.C., Amal, R., 1998. Fractal aggregates of polydisperse Letterman, R.D., Amirtharajah, A., O’Melia, C.R., 1999.
particles. Journal of Colloid and Interface Science 205, Coagulation and flocculation. In: Letterman, R.D. (Ed.), Water
459e469. Quality and Treatment, fifth ed. AWWA, Denver, Colorado.
Bushell, G.C., Yan, Y.D., Woodfield, D., Raper, J., Amal, R., 2002. On McGraw-Hill Inc.; New-York.
techniques for the measurement of the mass fractal Li, H., Addai-Mensah, J., Thomas, J.C., Gerson, A.R., 2005. The
dimension of aggregates. Advances in Colloid and Interface influence of Al(III) supersaturation and NaOH concentration
Science 95, 1e50. on the rate of crystallization of Al(OH)3 precursor particles
Cañizares, P., Jiménez, C., Martiñez, F., Rodrigo, M.A., Sáez, C., from sodium aluminate solutions. Journal of Colloid and
2009. The pH as a key parameter in the choice between Interface Science 286, 511e519.
coagulation and electrocoagulation for treatment of Mandelbrot, B.B., 1982. The Fractal Geometry of Nature. W. H.
wastewater. Journal of Hazardous Materials 163, 158e164. Freeman and Co, New-York.
Chakraborti, R.K., Gardner, K.H., Atkinson, J.F., Van Matsui, Y., Yuasa, A., Furuya, Y., Kamei, T., 1998. Dynamic
Benschoten, J.E., 2003. Changes in fractal dimension during analysis of coagulation with Alum and PACl. Journal of
aggregation. Water Research 37, 873e883. American Water Works Association 90, 96e106.
Chen, G., Chen, X., Yue, P.L., 2000. Electrocoagulation and Matteson, M.J., Dobson, R.L., Glenn Jr., R.W., Kukunoor, N.S.,
electroflotation of restaurant wastewater. Journal of Waits III., W.H., Clayfield, E.J., 1995. Electrocoagulation and
Environmental Engineering 126 (9), 858e863. separation of aqueous suspensions of ultrafine particles.
Duan, J., Gregory, J., 2003. Coagulation by hydrolyzing salts. Colloids and Surfaces A: Physicochemical Engineering Aspects
Advances in Colloid and Interface Science 100e102, 475e502. 104, 101e109.
Flesch, J.C., Spicer, P.T., Pratsinis, S.E., 1999. Laminar and Mollah, M.Y., Schennech, R., Parga, J.R., Cocke, D.L., 2001.
turbulent shear-induced flocculation of fractal aggregates. Electrocoagulation (EC) e science and applications. Journal of
AIChE Journal 45 (5), 1114e1124. Hazardous Materials B 84, 29e41.
Guan, J., Waite, T.D., Amal, R., 1998. Rapid structure Mollah, M.Y., Morkovsky, P.G., Gomes, A.G., Kesmez, M., Parga, J.,
characterization of bacterial aggregates. Environmental Cocke, D.L., 2004. Fundamentals, present and future
Science and Technology 32, 3735e3742. perspectives of electrocoagulation. Journal of Hazardous
Harif, T., Adin, A., 2007. Characteristics of aggregates formed by Materials B114, 199e210.
electroflocculation of a colloidal suspension. Water Research Ogutveren, U.B., Koparal, S., 1992. Electrocoagulation for
41, 2951e2961. oilewater emulsion treatment. Journal of Environmental
Harif, T., Adin, A., 2011. Size and structure evolution of kaolin- Science and Health, Part A: Toxic/Hazardous Substances and
Al(OH)3 flocs in the electroflocculation process: a study Environmental Engineering 32 (9e10), 2507e2520.
using static light scattering. Water Research 45 (18), Oles, V., 1992. Shear-induced aggregation and breakup of
6195e6206. polystyrene latex particles. Journal of Colloid and Interface
Harif, T., Hai, M., Adin, A., 2006. Electroflocculation as potential Science 154, 351e358.
pretreatment in colloid ultrafiltration. Water Science and Picard, T., Cathalifaud-Feuillade, G., Mazet, M.,
Technology 6 (1), 69e78. Vandensteendam, C., 2000. Cathodic dissolution in the
Hogg, R., 2005. Flocculation and dewatering of fine-suspension electrocoagulation process using aluminum electrodes.
particles. In: Stechemesser, H., Dobiás, B. (Eds.), Coagulation Journal of Environmental Monitoring 2, 77e80.
and Flocculation, second ed. CRC Press, Florida. Potanin, A.A., 1991. On the mechanism of aggregation and
Holt, P.K., Barton, G.W., Wark, M., Mitchell, C.A., 2002. A breakup of polystyrene. Journal of Colloid and Interface
quantitative comparison between chemical dosing and Science 145, 140e157.
electrocoagulation. Colloids and Surfaces A: Physicochemical Pouet, M.F., Grasmick, A., 1995. Urban wastewater treatment by
Engineering Aspects 211, 233e248. electrocoagulation and flotation. Water Science and
Ivanshivili, A.I., Przhegorlinskii, V.I., Kalichenko, T.D., 1987. Technology 31 (3e4), 275e283.
Comparative evaluation of the efficiency of electrocoagulation Shahjahan Kaisar Alam Sarkar, Md., Evans, G.M., Donne, S.W.,
and reagent methods of clarifying wastewater. Soviet Journal 2010. Bubble size measurement in electroflotation. Minerals
of Water Chemistry and Technology 9 (5), 118e119. Engineering 23, 1058e1065.
Jarvis, P., Jefferson, B., Gregory, J., Parsons, S.A., 2005a. A review of Sorensen, C.M., 1997. Scattering and absorption of light by
floc strength and breakage. Water Research 39 (14), 3121e3137. particles and aggregates. In: Birdi, K.S. (Ed.), Handbook of
Jarvis, P., Jefferson, B., Parsons, S.A., 2005b. Measuring of floc Surface and Colloid Chemistry, first ed. CRC Press Ltd., Florida.
structural characteristics. Reviews in Environmental Science Spicer, P.T., Pratsinis, S.E., 1996. Shear-induced flocculation: the
and Bio/Technology 4 (1e2), 1e18. evolution of floc structure and the shape of the size
Jarvis, P., Jefferson, B., Parsons, S.A., 2006. Floc structural distribution at steady-state. Water Research 30 (5),
characteristics using conventional coagulation for a high doc, 1049e1056.
low alkalinity surface water source. Water Research 40 (14), Spicer, P.T., Pratsinis, S.E., Raper, J., Amal, R., Bushell, G.,
2727e2737. Meesters, G., 1998. Effect of shear schedule on particle size,
Kusters, K.A., Wijers, J.G., Thoenes, D., 1997. Aggregation kinetics density, and structure during flocculation in stirred tanks.
of small particles in agitated vessels. Chemical Engineering Powder Technology 97, 26e34.
Science 52 (1), 107e121. Sposito, G., 1996. The Environmental Chemistry of Aluminum,
Lakshmanan, D., Clifford, D.A., Samanta, G., 2009. Ferrous and second ed. CRC Press Ltd., Florida.
ferric ion generation during iron electrocoagulation. Tambo, N., 1991. Basic concepts and innovative turn of
Environmental Science and Technology 43, 3853e3859. coagulation/flocculation. Water Supply 9, 1e10.
Larue, O., Vorobiev, E., Vu, C., Durand, B., 2003. Electrocoagulation Tang, S., Preece, J.M., McFarlane, C.M., Zhang, Z., 2000.
and coagulation by iron of latex particles in aqueous Fractal morphology and breakage of DLCA and RLCA
suspensions. Separation and Purification Technology 31, aggregates. Journal of Colloid and Interface Science 221,
177e192. 114e123.
3188 w a t e r r e s e a r c h 4 6 ( 2 0 1 2 ) 3 1 7 7 e3 1 8 8

Thill, A., Lambert, S., Moustier, S., Ginestet, P., Audic, J.M., Waite, T.D., 1999. Measurement and implications of floc structure
Bottero, J.Y., 2000. Structural interpretations of static in water and wastewater treatment. Colloids and Surfaces A:
light scattering patterns of aggregates II. Experimental Physicochemical Engineering Aspects 151, 27e41.
study. Journal of Colloid and Interface Science 228, Zhu, B., Clifford, D.A., Shankararaman, C., 2005. Comparison
386e392. of electrocoagulation and chemical coagulation
Vik, E.A., Carlson, D.A., Eikum, A.S., Gjessing, E.T., 1984. pretreatment for enhanced virus removal using
Electrocoagulation of potable water. Water Research 18 (11), microfiltration membranes. Water Research 39 (13),
1355e1360. 3098e3108.

You might also like