You are on page 1of 8

w a t e r r e s e a r c h 6 8 ( 2 0 1 5 ) 1 1 9 e1 2 6

Available online at www.sciencedirect.com

ScienceDirect

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

Dependence of floc properties on coagulant type,


dosing mode and nature of particles

Wenzheng Yu a,b, John Gregory a,*, Luiza C. Campos a, Nigel Graham b


a
Department of Civil, Environmental and Geomatic Engineering, University College London, Gower Street, London
WC1E 6BT, UK
b
Department of Civil and Environmental Engineering, Imperial College London, South Kensington Campus, London
SW7 2AZ, UK

article info abstract

Article history: Kaolin suspensions were coagulated with AlCl3 and a high-basicity PACl at pH 7, at dosages
Received 26 June 2014 that gave zeta potentials close to zero. The actions of the two coagulants were completely
Received in revised form different. With AlCl3, the formation of an amorphous hydroxide precipitate played a
23 September 2014 dominant role. When the coagulant was added to the suspension, flocs grew rapidly and
Accepted 25 September 2014 incorporated most of the kaolin particles within the hydroxide precipitate. When the
Available online 7 October 2014 suspension was added some time after the coagulant, the clay particles were found to be
mainly on the outer floc surfaces, although the floc size was about the same. The light
Keywords: scattering properties of the flocs were very dependent on the number and location of
Coagulation particles in the precipitate. With PACl, delaying the addition of kaolin had no influence on
Al(III) the final floc properties.
PACl In further tests, different suspensions over a range of concentrations were coagulated
Hydroxide precipitate with alum at pH 7. Monitoring by a turbidity fluctuation technique showed an apparent
Turbidity fluctuations increase in floc size with increasing particle concentration. However, floc sizes determined
from microscope images were very nearly constant, independent of particle nature and
concentration. With different particle types, the monitoring results were greatly dependent
on the light scattering properties of the particles.
Particles incorporated within hydroxide flocs appeared to have no influence on floc
properties, such as size and strength.
2014 Elsevier Ltd. All rights reserved.

been investigated in many previous studies (Dempsey et al.,


1. Introduction 1984; Gray et al., 1995; Matsui et al., 1998; Xiao et al., 2008;
Hussain et al., 2013).
Coagulation is a widely-used process for conventional water Al(III) and PACl have different coagulation mechanisms
treatment, and Al(III) is a very common coagulant in treating and efficiencies. PACl has often been found to be more effi-
drinking water. Pre-hydrolyzed coagulants such as poly- cient in the removal of natural organic matter (NOM) from
aluminum chloride (PACl) have been known for many years high alkalinity and high pH water than traditional coagulants
(e.g. Bottero et al., 1980) and their use in water treatment has based on Al(III) and Fe(III) (Yan et al., 2008a) or in neutral pH

* Corresponding author. Tel.: 44 20 7689 7818.


E-mail addresses: j.gregory@ucl.ac.uk, johngregory15@hotmail.com (J. Gregory).
http://dx.doi.org/10.1016/j.watres.2014.09.045
0043-1354/ 2014 Elsevier Ltd. All rights reserved.
120 w a t e r r e s e a r c h 6 8 ( 2 0 1 5 ) 1 1 9 e1 2 6

water through charge neutralization and bridging (Yan et al., coagulation dominates the coagulation mechanism, a small
2008b), especially at low dosage (Zouboulis et al., 2008). PACl additional dosage of alum added at the time of floc breakage,
proved to be a more efficient coagulant than alum and pro- can greatly improve the re-growth ability of broken flocs.
duced treated water with low turbidity and residual However a similar small addition of high-basicity PACl has no
aluminum content without requiring acid addition for pH effect on floc re-growth, indicating that floc surfaces are quite
adjustment and subsequent base addition for re-stabilizing different in the two cases.
the water (Zouboulis et al., 2008; Zarchi et al., 2013). PACls In this work, coagulation of clay suspensions with Al(III)
with higher basicity (OH/Al) are more efficient for turbidity coagulants and a high-basicity PACl has been carried out with
and NOM removal (Yan et al., 2008b,c). PACls with a high different dosing conditions and with different suspensions.
content of the Al tridecamer (Al13) can give larger flocs, higher The main aims of the work were:
floc growth rate and shorter floc settling times than other Al-
based coagulants (Hussain et al., 2013). However, some re-  To compare the effects of Al(III) and PACl coagulants on the
searchers have come to different conclusions. Trinh and Kang structure of flocs, by optical microscopy
(2011) found that PACl was more effective in removing parti-  To gain further insight into the mechanism of floc forma-
cles from water, but less effective in removal of NOM tion by adding particles at different times after coagulant
compared with alum. Shi et al. (2007) also found that co- addition
agulants with preformed Al species were less effective than  To investigate the effect of the nature and number of par-
conventional Al coagulant in removing humic acid with high ticles included within hydroxide flocs on the apparent floc
molecular weight and hydrophobic character, giving smaller size determined by an optical technique
flocs.
The coagulation mechanisms of alum and PACl are
different. Under most practical conditions, precipitate for-
mation and sweep flocculation occur with alum, whereas 2. Materials and methods
charge neutralization by the Al13 polycation, electrostatic
patch and bridging interactions are more important for PACl 2.1. Materials
(Wu et al., 2009). After dosing into water, Al(III) undergoes a
series of hydrolysis, polymerization, precipitation and aggre- Kaolin clay (Imerys, St Austell, Cornwall, UK) was used as a
gation processes. Amorphous Al(OH)3 has a low solubility model suspension. 200 g of kaolin was dispersed in 500 mL of
around neutral pH and can remove impurities by a combina- deionized water in a high-speed blender. To obtain full
tion of adsorption and precipitate enmeshment. PACl also dispersion it was necessary to raise the pH of the suspension
forms precipitates at sufficiently high pH, but the polymeric to about 7.5, which was achieved by adding 5 mL of 0.1 mM
structure of PACl is retained upon precipitation yielding a NaOH. After blending at 4000 rpm for 10 min the clay sus-
solid phase with different light scattering characteristics, pension was diluted to 1 L with deionized water and allowed
electrophoretic mobility and solubility than amorphous to stand overnight (12 h) in a measuring cylinder. The top
Al(OH)3 (Van Benschoten and Edzwald, 1990). Klute (1990) 800 mL was decanted and its solids content was determined
found that the preformed species in PACls are more stable gravimetrically and found to be 133 g/L.
than those formed in situ after alum addition; Matsui et al. Spherical silica particles (Geltech Inc) were dispersed in
(1998) assumed that those preformed species could destabi- dilute NaOH solution to give a concentration of 10 g/L. These
lize particles faster. The preformed polymeric species con- particles were porous, with a density of 2 g/cm3 and a mean
tained in PACl are fairly stable and highly positively charged diameter of 1.5 mm, with a narrow size distribution.
and these species can be adsorbed on surfaces of clay and Polystyrene latex was prepared by emulsifier-free poly-
other particles and undergo surface aggregation and rear- merization of styrene, with potassium persulfate as initiator.
rangement to form electrostatic patches (Ye et al., 2007). The The particles were highly uniform with a mean diameter of
Al13 polycation ([Al13O4(OH)24(H2O)12]7) has a tetrahedral 1.6 mm, determined by an Elzone particle counter. The sample
Al(O)4 center surrounded by 12 octahedrally coordinated Al had been purified by repeated sedimentation, decantation and
atoms (Johansson, 1960). Further hydrolysis or rearrangement redispersion in deionized water and the final concentration
of Al13 subunits can occur (Bottero et al., 1987; Bertsch, 1987). was 40 g/L.
Both Al(OH)3-rich and Al13-aggregate flocs do not possess AlCl3 and PACl25 (OH/Al ratio 2.5) were used as co-
long-range crystalline structure, but Al13-like crystals have agulants. AlCl3 was purchased from Aldrich. The PACl25 was
been found in the flocs formed by PACl and the chemical prepared using a laboratory base titration method at room
compositions of the Al(OH)3 precipitates and Al13 aggregates temperature (Wang et al., 2004) as described in detail in our
are markedly different (Lin et al., 2009). previous paper (Yu et al., 2011a). The final concentration of
In recent years, some researchers have focused on the floc both coagulants is 0.1 M (as Al). The samples, after being left to
breakage and re-growth process to better understand floc age one week, were analyzed using the ferron method as
characteristics and coagulation mechanisms. It has been described by Wang et al. (2004). AlCl3 was found to contain
found that only limited re-growth of broken flocs occurred for 97.1% Ala (mainly monomer Al) and PACl25 93.6% Alb (mainly
alum and PACl, indicating a significant irreversibility of the Al13).
floc break-up process (Yukselen and Gregory, 2004). It is likely Aluminum sulfate, Al2(SO4)3, NaHCO3 and HCl were of
that there are some changes on the surface of broken flocs. analytical reagent grade and prepared as a 0.1 M solutions in
According to our previous research (Yu et al., 2011b), if sweep deionized water.
w a t e r r e s e a r c h 6 8 ( 2 0 1 5 ) 1 1 9 e1 2 6 121

2.2. Jar test As flocs grow, their number concentration decreases, but
the increasing scattering cross section more than compen-
For the most of the flocculation tests, the stock kaolin sus- sates, so that the FI value increases.
pension was diluted in water (800 mL) to give a clay concen- Of course, in a real flocculating suspension, flocs will have
tration of 50 mg/L. The flocculation tests were carried out in a range of sizes, and hence a range of scattering cross sections.
5 mM NaHCO3 solutions, made up with deionized (DI) water. In that case, C can be regarded as an average scattering cross
During the jar test, water pH was maintained at 7.0 via prior section, but larger flocs have a greater influence on the
addition of a pre-determined amount of 0.1 M HCl. The dosage average value.
of AlCl3 and PACl25 was 0.1 mM and 0.02 mM respectively (as When clay suspensions are dosed with hydrolyzing co-
Al), since the zeta potential of the flocs was near zero at these agulants, the resulting flocs are composed of amorphous
dosages (see below). The temperature throughout was hydroxide precipitate with included clay particles. In such
25 1  C. The jar tests were carried out with a Flocculator 2000 cases, the clay particles can give the dominant contribution
(Kemira Kemi, Helsingborg, Sweden), which enables stirring to the optical properties of the flocs and the scattering cross
speeds and times to be set in advance. section may be considerably increased, even though the
In order to investigate the effect of delaying the addition of actual floc size may be little changed (Gregory, 2009b).
kaolin suspension, a fixed dosage of alum or PACl25 was added
into the test solution and at the same time the stirring of 2.4. Floc images and electrophoretic mobility
flocculator started. The stirring speed was set at 200 rpm
(G z 184 s1) for 60 s, to give rapid mixing and then 50 rpm During the slow stirring period, samples of flocs were taken
(23 s1) to allow floc growth. Kaolin suspension (50 mg/L) was from below the surface of the suspension with a hollow glass
added either initially (i.e. before rapid mixing) or at 0.5, 1, 2, 5 tube with an inner diameter of 5 mm. One end of tube was
or 10 min after coagulant addition. After kaolin addition, inserted 3.0 cm below the surface and the other end was
50 rpm stirring was maintained for a further 15 min. covered by a finger, and then the sample was carefully with-
The effects of different types of particle at different con- drawn. After transferring the sample onto a flat microscope
centrations were investigated by a similar jar test procedure. slide, an image of the flocs was captured by an optical mi-
However, in this case, the flocs were broken at high shear croscope with a CCD camera (GE-5, Aigo, China). The images
(200 rpm for 1 min) and then allowed to re-grow at 50 rpm. In were analyzed by Image J software and floc diameters were
these tests alum at 0.1 mM Al was the coagulant and the pH determined from the floc area.
was 7. Electrophoretic mobility (EM) was measured on the kaolin
suspension before coagulant addition and for samples after
coagulant dosing and 1 min of rapid mixing, by a Particle
2.3. PDA measurement
Electrophoresis Apparatus Mk 2 (Rank Brothers Ltd, Cam-
bridge, UK). In addition, the EM of flocs after 1 min of breakage
Experiments on the kinetics of formation, breakage and sub-
at high stirring speed was measured. The delay time of the
sequent re-growth of flocs were performed using the
measurement was about 40 s. The average EM value for a
Turbidity Fluctuation method (PDA3000, Rank Brothers, UK).
sample was determined from 20 measurements.
The procedure was similar to that of Yukselen and Gregory
(2004). The PDA 3000 measures the average transmitted light
intensity (dc value) and the root mean square (rms) value of
the fluctuating component of the intensity. The ratio (rms/dc) 3. Result and discussion
provides a sensitive measure of particle aggregation. The ratio
value is strongly correlated with floc size and always increases 3.1. Electrophoretic mobility of flocs
as flocs grow larger. It is referred to here as the Flocculation
Index (FI), which significantly increases as aggregation occurs, AlCl3 and PACl25, at a given dosage, cause different EM of flocs.
and decreases when aggregates are broken. After the FI value Prior to dosage, the EM of kaolin particles was negative
reached an initial steady value, coagulant was added to the (around 3.9 mm s1 V1 cm1). The addition of low dosages of
suspension and the dc and rms values recorded every second. AlCl3 caused a rapid increase of EM and the charge became
FI values depend essentially on the number concentration nearly stable at the dosage of 0.1 mM Al. Higher dosage caused
of particles (flocs) and their extinction (or scattering) cross only a slight further increase of EM. For PACl25, very low
sections (Gregory, 2009a). The FI value always increases as dosages caused a steep rise in EM and charge reversal
particles aggregate and decreases when aggregates are occurred at about 0.02 mM Al. Breakage of flocs caused no
broken. It depends essentially on the number concentration, significant change in EM. The dosages of PACl25, and AlCl3
N, and effective light scattering cross-section, C, of aggregates required to give EM z 0 (the isoelectric point) were 0.02 mM
(flocs). For quite large flocs, the scattering cross section should and 0.1 mM, respectively. In the following results, just these
be proportional to the geometric cross-sectional area of a floc. two dosages were used (Fig. 1).
If all the flocs are of equal size then the FI value is given by:
1 = 3.2. Effect of delayed addition of kaolin suspension
FI KCN 2
(1)

where K is a constant that depends only on the instrument The effect of different delay times with AlCl3 and PACl25 on the
(essentially on the dimensions of the light beam). apparent coagulation efficiency was explored by adding kaolin
122 w a t e r r e s e a r c h 6 8 ( 2 0 1 5 ) 1 1 9 e1 2 6

unaffected by the delay. With AlCl3 at pH 7, precipitation to


give an amorphous precipitate occurs during the delay time
and this markedly reduces the coagulation efficiency.

3.3. Images of flocs

In order to see any changes in floc characteristics when the


kaolin was added later than coagulants, images of flocs were
captured after flocculation. Usually, after AlCl3 is added into
the kaolin suspension at pH 7, kaolin particles will adhere to
the colloidal precipitate particles, which then aggregate to
form larger flocs. These flocs essentially consist of amorphous
precipitate with embedded kaolin particles, distributed
randomly throughout the floc (sweep floc). This idea is
confirmed by the micrograph in Fig. 3a, where the clay parti-
cles can be seen within the amorphous hydroxide precipitate.
By contrast, when kaolin was added 10 min after addition
Fig. 1 e Variation of electrophoretic mobility of kaolin flocs
of AlCl3, the appearance of flocs was very different (Fig. 3b).
with coagulant dosage at pH 7.
The size of flocs in Fig. 3b is nearly the same as those in Fig. 3a,
but the kaolin particles were distributed in a very different
way. In Fig. 3b, most of the particles are located on the outside
of the flocs, with very few in the interior. Also each floc in
particles at different times after coagulant addition and the Fig. 3b had fewer kaolin particles than those in Fig. 3a. During
results are shown in Fig. 2. the 10 min after coagulant addition hydroxide precipitates are
The results were significantly different for AlCl3 and formed and these grow to large flocs. Kaolin particles added
PACl25. From Fig. 2a, when the kaolin and AlCl3 were added at subsequently have difficulty in entering these flocs and
the same time, the FI value was the highest, and the final adhere mainly to the outer floc surface. Comparing Fig. 2a
plateau FI value decreased markedly as the delay time with Fig. 3a and b, it seems that the FI value was not deter-
increased. However, we cannot assume that the lower FI value mined by the size of flocs, but by their kaolin content. The very
implies smaller flocs, because of different light scattering small increase in FI before kaolin addition is because hy-
properties of the flocs (see 3.3). droxide flocs scatter little light. Only when kaolin is added
With PACl25 as the coagulant, the results were very does the FI value rise significantly.
different. As shown in Fig. 2b, the rate of increase of FI did not The images of flocs produced by PACl25 (Fig. 3c and d) are
change as the delay time increased. Also the plateau FI values completely different from those for AlCl3. Firstly, the PACl
for different delay times were nearly the same, whereas for flocs are considerably smaller and denser. Also it is clear that
AlCl3, the coagulation rate decreased with delay time. It can be there is no significant effect of delaying the addition of kaolin
assumed that PACl25, in solution at pH 7, does not change suspension. These observations are consistent with a charge-
significantly during the delay time because of its high positive neutralization mechanism for PACl coagulation, either by Al13
charge. Thus, its ability to coagulate clay particles remains polycations or small aggregates of these. It may be that

Fig. 2 e The effect of delay time before adding kaolin: (a) 0.1 mM AlCl3, (b) 0.02 mM PACl25. AlCl3 or PACl25 was added into the
test solution (800 mL) and at the same time the stirring of flocculator started. The stirring speed was set at 200 rpm for 60 s
and then at 50 rpm. Kaolin suspension was added at 0, 0.5, 1, 2, 5 and 10 min, after coagulant addition. Note that the 0.5 and
1 min additions were during the rapid mix (200 rpm) phase.
w a t e r r e s e a r c h 6 8 ( 2 0 1 5 ) 1 1 9 e1 2 6 123

Fig. 3 e Images of flocs with different delay times before the addition of kaolin particles: (a) AlCl3 with no delay, (b) AlCl3 with
10 min delay, (c) PACl25 with no delay, (d) PACl25 with 10 min delay.

polycation aggregates are able to adsorb in a patchwise flocs, so that, despite their smaller size, they give about the
manner and bind particles by electrostatic attraction or by a same FI value.
bridging mechanism. It seems that the nature of the coagulant
species formed by PACl25 does not change significantly during 3.4. Effect of different particle concentrations
the 10 min delay time and there is no apparent difference in
their behavior as coagulants. Suspensions of, kaolin, silica and polystyrene latex at
The flocculation results in Fig. 2 can be explained by the different concentrations (2e10 mg/L) were coagulated by alum
images in Fig. 3. In the case of AlCl3, with no delay before at a dosage of 0.1 mM Al and at pH 7. The results are shown in
adding kaolin, the FI value reaches about 1.4 and the corre- Fig. 4. In all cases, the FI value increased during the floccula-
sponding value for PACl25 is only slightly less, despite the tion process, but to very different extents, depending on the
great difference in floc size shown in Fig. 3. Eq (1) shows that particle concentration. As the concentration of kaolin
the FI value depends inversely on the number concentration, increased from 2 mg/L to 10 mg/L, the first plateau of the FI
N, and directly on the light scattering cross section, C, of value increased as well as the corresponding value for the re-
particles or flocs. In the case of amorphous hydroxide flocs, grown flocs. With the other particles, similar trends were
without included particles, C is very low, so that, even for large observed. The right-hand graphs in Fig. 4 show plateau FI
flocs (and hence low N) the FI value does not increase greatly values plotted against particle concentration. In all cases good
until kaolin is added. However, when kaolin particles are linear correlations were found for the flocs before and after
included in the amorphous flocs, they can greatly increase the breakage.
effective value of the scattering cross section. So, the scat- There are differences between FI values for different
tering power of a hydroxide floc depends largely on the particles at equivalent concentrations. For the flocs before
included particles and different FI values reflect different breakage, the slopes of the regression lines in Fig. 4 show that
particle content and not necessarily floc size (Gregory, 2009b). silica particles give about 25% higher FI values than kaolin,
In Fig. 2a, the decrease of FI value with increasing delay time and that latex particles show a 7-fold increase over kaolin.
can be explained in this way. It is likely that the flocs have Some of this effect can be explained by the different densities
about the same size, independent of delay time, but the of the particles. These are about 2.7, 2.0 and 1.05 g/cm3 for
decreasing FI values are due to the smaller number of kaolin kaolin, silica and latex respectively. The particle concentra-
particles in each floc. Similar effects with other particles are tions in Fig. 4 are expressed on a mass basis, but the volume
shown in the next section. concentrations differ and these are more relevant to light
The flocs produced by PACl25 (Fig. 3c and d) are more dense scattering and turbidity. For the same mass concentration,
and have higher scattering cross-sections than the hydroxide the volume fractions of kaolin, silica and latex would be in
124 w a t e r r e s e a r c h 6 8 ( 2 0 1 5 ) 1 1 9 e1 2 6

Fig. 4 e Formation breakage and re-growth of alum flocs at pH 7, with different particles at different concentrations. The left-
hand plots show the variation of FI value with time. The corresponding maximum FI values before floc breakage and after
re-growth are shown on the right-hand side. (a, b) kaolin, (c, d) silica, (e, f) latex.

the ratio 1:1.3:2.6. So, for silica, the increase in FI value over sizes derived as before. At least 100 flocs were sampled in each
kaolin is nearly the same as the volume fraction increase. case and the average sizes and standard deviations are shown
However, for latex, the higher volume fraction accounts for in Table 1. For low concentrations of kaolin and silica, the flocs
less than half of the increase in FI value, indicating that the were quite difficult to resolve by light microscope and some
polystyrene particles are more optically dense than kaolin values are omitted from the Table. All of the floc sizes in Table
and silica. 1 are close to 170 mm, despite the very different FI values
Images of flocs formed for the different particle concen- in Fig. 4. This confirms the idea that the FI value does not give
trations, after breakage and re-growth, were analyzed and floc an absolute indication of size for hydroxide flocs, since it
w a t e r r e s e a r c h 6 8 ( 2 0 1 5 ) 1 1 9 e1 2 6 125

Table 1 e Average floc size after re-growth with different Acknowledgments


particles at different concentrations. Coagulation with
alum (0.1 mM Al) at pH 7. Standard deviations given in
This work was supported by National Natural Science Foun-
brackets. (ND: not determined).
dation of China (Grants 51108444). The authors wish to
Particles Floc size (mm)
acknowledge the support of the European Commission in
mg/L
2 4 6 8 10 providing a Marie Curie International Incoming Fellowship for
Kaolin ND ND 174 (65) 170 (62) 171 (62) Dr Wenzheng Yu.
Silica ND 174 (67) 172 (62) 168 (61) 165 (58)
Latex 168 (68) 167 (70) 164 (62) 161 (63) 165 (74)

references

depends greatly on the nature and number of particles Bertsch, P.M., 1987. Conditions for Al-13 polymer formation in
included within the flocs. Nevertheless, results such as those partially neutralized aluminum solutions. Soil Sci. Soc. Am. J.
in Figs. 2 and 4 can be very useful in deriving relative infor- 51, 825e828.
Bottero, J.Y., Cases, J.M., Fiessinger, F., Poirier, J.E., 1980. Studies of
mation on flocculation rate, floc breakage and re-growth.
hydrolyzed aluminum-chloride solutions .1. Nature of
It is also noteworthy that the presence of different par-
aluminum species and composition of aqueous-solutions. J.
ticles within hydroxide flocs has no significant effect on floc Phys. Chem. 84, 2933e2939.
properties, such as strength, breakage and re-growth. The Bottero, J.Y., Axelos, M., Tchoubar, D., Cases, J.M., Fripiat, J.J.,
plateau floc size under given shear conditions is a good Fiessinger, F., 1987. Mechanism of formation of aluminum
indication of floc strength. This is not affected by included trihydroxide from keggin Al 13 polymers. J. Colloid Interface
particles, over the concentration range 2e10 mg/L. The floc Sci. 117, 47e57.
Dempsey, B.A., Ganho, R.M., O'Melia, C.R., 1984. The coagulation
properties are determined by the nature of the hydroxide
of humic substances by means of aluminum salts. J. Am.
precipitate. Water Works Assoc. 76, 141e150.
Gray, K.A., Yao, C.H., O'Melia, C.R., 1995. Inorganic metal
polymers - preparation and characterization. J. Am. Water
Works Assoc. 87, 136e146.
4. Conclusions Gregory, J., 2009a. Monitoring particle aggregation processes. Adv.
Colloid Interface Sci. 147-48, 109e123.
Gregory, J., 2009b. Optical monitoring of particle aggregates. J.
The main conclusions of this work are: Environ. Sci. 21, 2e7.
Hussain, S., van Leeuwen, J., Chow, C., Beecham, S.,
1. With AlCl3 as coagulant at pH 7, hydroxide precipitation Kamruzzaman, M., Wang, D.S., Drikas, M., Aryal, R., 2013.
plays the dominant role. If kaolin suspension is added Removal of organic contaminants from river and reservoir
some time after the coagulant, when some precipitation waters by three different aluminum-based metal salts:
coagulation adsorption and kinetics studies. Chem. Eng. J. 225,
has occurred, then the resulting flocs are very different in
394e405.
structure from those formed with addition of coagulant to
Johansson, G., 1960. On the crystal structures of some basic
a kaolin suspension. In the latter case, clay particles are aluminium salts. Acta Chem. Scand. 14, 771e773.
distributed throughout the hydroxide floc, whereas with a Klute, R., 1990. Destabilization and aggregation in turbulent pipe
delay time particles are found mostly on the outer floc flow. In: Hahn, H.H., Klute, R. (Eds.), Chemical Water and
surface and fewer particles are removed. For PACl25, Wastewater Treatment. Springer-Verlag, Berlin, pp. 33e54.
delaying the addition of kaolin had no significant effect on Lin, J.L., Huang, C.P., Chin, C.J.M., Pan, J.R., 2009. The origin of
Al(OH)(3)-rich and Al-13-aggregate flocs composition in PACl
floc growth, because the coagulation mechanism was quite
coagulation. Water Res. 43, 4285e4295.
different, being mainly based on charge neutralization. Matsui, Y., Yuasa, A., Furuya, Y., Kamei, T., 1998. Dynamic
2. With alum at pH 7 and different suspension concentra- analysis of coagulation with alum and PACl. J. Am. Water
tions, the response of a turbidity fluctuation monitor de- Works Assoc. 90, 96e106.
pends directly on the included particles within hydroxide Shi, B.Y., Wei, Q.S., Wang, D.S., Zhu, Z., Tang, H.X., 2007.
flocs. Over a certain concentration range, there is a linear Coagulation of humic acid: the performance of preformed and
relationship between Flocculation Index and the particle non-preformed Al species. Colloids Surf. A Physicochem. Eng.
Aspects 296, 141e148.
concentration. However, floc size is independent of particle
Trinh, T.K., Kang, L.S., 2011. Response surface
concentration. The FI value is also influenced by the nature methodological approach to optimize the coagulation-
of the particles. Those that scatter more light give higher FI flocculation process in drinking water treatment. Chem. Eng.
values. Res. Des. 89, 1126e1135.
3. The strength of flocs produced by sweep coagulation is Van Benschoten, J.E., Edzwald, J.K., 1990. Chemical aspects of
almost entirely determined by the nature of the hydroxide coagulation using aluminum salts .1. Hydrolytic reactions of
alum and polyaluminum chloride. Water Res. 24, 1519e1526.
precipitate and not by the included particles.
Wang, D.S., Sun, W., Xu, Y., Tang, H.X., Gregory, J., 2004.
4. The light scattering properties of hydroxide flocs depend
Speciation stability of inorganic polymer flocculant-PACl.
mostly on the included particles and this has important Colloids Surf. A Physicochem. Eng. Aspects 243, 1e10.
implications for monitoring of floc size by light scattering Wu, X.H., Ge, X.P., Wang, D.S., Tang, H.X., 2009. Distinct
methods. mechanisms of particle aggregation induced by alum and
126 w a t e r r e s e a r c h 6 8 ( 2 0 1 5 ) 1 1 9 e1 2 6

PACl: floc structure and DLVO evaluation. Colloids Surf. A polyaluminum chlorides: role of electrostatic patch. Colloids
Physicochem. Eng. Aspects 347, 56e63. Surf. A-Physicochem. Eng. Aspects 294, 163e173.
Xiao, F., Ma, J., Yi, P., Huang, J.C.H., 2008. Effects of low Yu, W.Z., Gregory, J., Campos, L., Li, G.B., 2011a. The role of mixing
temperature on coagulation of kaolinite suspensions. Water conditions on floc growth, breakage and re-growth. Chem.
Res. 42, 2983e2992. Eng. J. 171, 425e430.
Yan, M.Q., Wang, D.S., Qu, J.H., Ni, J.R., Chow, C.W.K., 2008a. Yu, W.Z., Gregory, J., Campos, L.C., 2011b. Breakage and re-growth
Enhanced coagulation for high alkalinity and micro-polluted of flocs: effect of additional doses of coagulant species. Water
water: the third way through coagulant optimization. Water Res. 45, 6718e6724.
Res. 42, 2278e2286. Yukselen, M.A., Gregory, J., 2004. The reversibility of floc
Yan, M.Q., Wang, D.S., Ni, J.R., Qu, J.H., Chow, C.W.K., Liu, H.L., breakage. Int. J. Miner. Process. 73, 251e259.
2008b. Mechanism of natural organic matter removal by Zarchi, I., Friedler, E., Rebhun, M., 2013. Polyaluminium chloride
polyaluminum chloride: effect of coagulant particle size and as an alternative to alum for the direct filtration of drinking
hydrolysis kinetics. Water Res. 42, 3361e3370. water. Environ. Technol. 34, 1199e1209.
Yan, M.Q., Wang, D.S., Yu, J.F., Ni, J.R., Edwards, M., Qu, H.H., Zouboulis, A., Traskas, G., Samaras, P., 2008. Comparison of
2008c. Enhanced coagulation with polyaluminum chlorides: efficiency between poly-aluminium chloride and aluminium
role of pH/Alkalinity and speciation. Chemosphere 71, sulphate coagulants during full-scale experiments in a
1665e1673. drinking water treatment plant. Sep. Sci. Technol. 43,
Ye, C.Q., Wang, D.S., Shi, B.Y., Yu, J.F., Qu, J.H., Edwards, M., 1507e1519.
Tang, H.X., 2007. Alkalinity effect of coagulation with

You might also like