Sie sind auf Seite 1von 23

8) Pergamon Wal. Sci. Tech. Vol. 31, No. 3-4, pp. 1-23, 1995.

Copyright © 1995 IAWQ


Printed in Great Britain. All rights reserved.
0273-1223/95 $9'50 + 0'00
0273-1223(95)00200-6

PRINCIPLES AND APPLICATIONS OF


DISSOLVED AIR FLOTATION

James K. Edzwald
Department of Civil and Environmental Engineering, University ofMassachusetts,
Marston Hall, Amherst. MA 01003-5205. USA

ABSTRACT
Principles of dissolved air flotation (DAF) discussed include: bubble formation and size, bubble-particle
interactions, measures of supplied air, and modeling of the reaction and clarification zones of the flotation
tank. Favorable flotation conditions for bubble attachment or adhesion to particles requires a reduction in the
charge of particles and production of hydrophobic particles or hydrophobic spots on particle surfaces. A
conceptual model for the bubble-particle reaction zone based on the single collector collision efficiency is
summarized and discussed. An alternative modeling approach is considered. Clarification or separation zone
modeling is based on particle-bubble agglomerate rise velocities. The application of DAF in drinking water
treatment is addressed beginning with summaries of design and operating parameters for several countries.
DAF should not be considered as a separate process, but integrated into the design and operation of the
overall treatment plant This concept shows that flocculation abead of DAF has different requirements
regarding floc size and strength compared to sedimentation. The efficiency of DAF in removing particles and
reducing particle loads to filters needs to be integrated into DAF plant design. The impact on filtration
performance is illustrated. Finally, fundamental and applied research needs are addressed.

KEYWORDS

Applications; bubbles; coagulation; dissolved air flotation; drinking water; filtration; flocculation;
hydrophobicity; particles; principles.

INTRODUCTION

Flotation has its beginnings in the mineral or ore processing industry. Because of its origins and subsequent
use in other particle separation applications, flotation nomenclature can be confusing. Sometimes, flotation
is described in terms of the material being removed or separated e.g., mineral flotation, precipitate flotation,
colloid flotation, and ion flotation. Other times it is described in terms of the method of bubble formation
e.g., electroflotation, dispersed air flotation, and dissolved air flotation. Dissolved air flotation can include
vacuum or pressurized methods. Vacuum flotation is limited to a pressure change of 1 atm (101.3 kPa) and
has limited applications (e.g., wastewater sludge thickening). The emphasis in this paper is pressurized
dissolved air flotation which is simply referred to as dissolved air flotation (OAF).

Dissolved air flotation was recognized as a method of separating particles (mineral ores) in the early 1900s.
A US patent was issued in 1905 for a process using pressurized aeration followed by pressure release
(Sulman et al., 1905). Kitehener (1984) has referred to a 1907 attribution of H. Norris in which small air
bubbles were formed "by supersaturating water with air at several atmospheric pressures and then injecting it
through a nozzle into a tank ...". From these early beginnings OAF has found many applications including:
2 J.K.EDZWALD

mineral separation; clarification of paper mill wastewaters, refinery wastewaters, combined sewer and storm
waters, municipal wastewaters in tertiary treatment, and oxidation pond effluents; municipal and industrial
waste sludge thickening; recycled paper de-inking; and wastewater reclamation. For drinking water
clarification, DAF was first used in the 1960s in South Africa and Scandinavia (Longhurst and Graham,
1987; Worrel, 1991; Haarhoff and van Vuuren, 1993). It is now widely used in these countries as well as in
Belgium, The Netherlands, and the United Kingdom, and is also used in Asia and Australia. It is an
emerging technology in North America that will become more important because of existing and proposed
regulations that require filtration of surface waters and increased removal of protozoa cysts such as
Cryptosporidium and Giardia. There are currently about eight operating plants in the USA. with others
under study and design.

Scope and Goals

This paper is divided into three major parts. The first section is devoted to general principles regardless of
application. The goals are twofold: (I) to summarize DAF principles pertaining to bubble formation and
size, particle-bubble interactions, and measures of bubble quantities released; and, (2) to summarize DAF
modeling approaches and concepts. The second section deals with drinking water applications. It likewise
has two goals: (I) to summarize existing design and operating information; and, (2) to make a case for
examining DAF in drinking water treatment with an integrated approach. The final section briefly addresses
fundamental and applied research needs.

PRINCIPLES

Bubbles

Bubble/ormation and size. Small air bubbles (100 um or less) are formed by injection of pressurized recycle
water into a flotation tank using specially designed nozzles or needle valves. The process of bubble
formation involves two steps: nucleation and growth. In a supersaturated system of clean water the large
pressure difference across the nozzle produces bubble nuclei spontaneously according to the thermodynamic
principle of minimizing the free energy change. Assuming air is an ideal gas, the critical diameter of the
bubble nucleus (deb)for homogeneous nucleation is

deb = 4 aiM' (I)

a =the surface tension


M' = the pressure change across the nozzle.

Figure I shows the critical diameter of the bubble nucleus as a function of the pressure change. Smaller
nuclei are formed at higher pressure changes. In a heterogeneous system, minimization of the free energy
change is made easier by bubble formation occurring on particle nuclei or other surfaces containing
scratches or crevices. The nuclei grow into bubbles in the second step.

Measurements of bubble sizes for DAF systems indicate bubbles maintain a steady state size range of 10 to
100 urn (Takahashi et al., 1979; Zabel, 1984; De Rijk et al., 1994). A reasonable estimate of the average
bubble diameter is 40 um. The steady state size depends on the saturator pressure and injection flow rate
(Takahashi et al., 1979). The injection flow or recycle flow must provide a rapid pressure drop and be
sufficient to prevent backflow and bubble growth on pipe surfaces in the vicinity of the injection system.
Higher pressures produce smaller bubbles, but there is a diminishing return in reducing the bubble size.
Above 500 kPa, increasing the saturator pressure has a small effect on bubble size (Heinanen et al., 1992;
De Rijk et al., 1994). To ensure small bubbles, pressure differences (saturator gauge pressures) of 400 to 600
kPa (4 to 6 atms) are recommended. Additional bubble growth may occur as the bubbles rise in the flotation
tank due to a decrease in the hydrostatic pressure or by coalescence. Both of these have negligible effects on
the small bubbles formed in DAF systems (Takahashi et al., 1979).
Dissolved air flotation 3

16

14

12

10
e-
e-
'Q
C 8
,D

-e" 6

0
0 200 400 600 800 1000
Pressure Change, P g (KPa)

Fig. I. Critical diam eter of bubble nucleu s for homogeneous nucleation of air in water at W · C as a function of the
pressure chang e.

Figure 2 is adapted from Clift et al. (1978) and shows the bubble rise velocity as a function of bubble size.
Small bubbles of about 100 um or less found in OAF systems rise as rigid spheres under laminar flow
conditions and obey Stokes law. Larger bubbles have higher rise velocities and exist as ellipsoids (I to 10
mm) or spherical caps ( > 10 mm).

Sph erical
'Y 10 ·Cap
'"
-!!:
E
~
.Q
~

£.
'y 1
0
'OJ
i>
'"
.~
l:t:
...
:;:;
J:l 0.1
=
lXl

0.01 .......- .....................&.-........- -................&--................................. -~.........


10 100 1000 toeeo
Equivalent Bubble Diameter, db ijLm)
Fig. 2. Bubble rise veloci ty of air bubble in water at 20·C (for bubbles of 100 IUD or less, experimental data follows
Stokes law; larger bubbles based on experimental data in Clift et al. (1978».

Bubble-particle interactions . Kitehener and Gochin (1981) list three possible mechanisms for forming
aggregates of bubbles and particles: (I) entrapment of preformed bubbles in large floc structures (floc size
4 J. K. EDZWALD

scale much larger than bubble size scale); (2) growth of bubble nuclei formation on particles or within floes
and (3) particle collision and adhesion with preformed bubbles. Mechanism I is more important where
larger particles or floes (IOO's of urn) either already exist (e.g., thickening) or are formed rapidly by high
rates of flocculation involving concentrated suspensions (e.g., municipal and industrial wastewaters).
Mechanism 2 probably occurs to varying degrees in most applications, however, it is Mechanism 3 that is
most important and applicable. This is true given the time scale ( < I sec (Rykaart and Haarhoff, 1995)) for
the formation of air bubbles from supersaturated recycle water injected into the flotation tank with pressure
changes of 4 to 6 atm and given its many applications in treating dilute suspensions. This is not to say, that
all of the supersaturated air comes out of solution instantaneously. No doubt some air leaves solution slowly.
Also it is noted that heterogeneous nucleation will be a factor in bubble formation, especially in applications
using clarified water as the recycle water.

Classically, the contact angle between the adsorbed bubble and particle as illustrated in Figure 3a is used to
describe the flotation of particles by bubbles. Here, the contact angle must be finite and large enough such
that the work or energy of adhesion of water to the solid particle is less than the energy of cohesion of the
water. A larger contact angle indicates hydrophobicity and good adhesion. The magnitude of the contact
angle; however, depends on the size scale of the bubbles and particles. The classical view of flotation arose
from separation of mineral ores in mining where both the sizes of bubbles and particles are relatively large
(i.e., dispersed air flotation, 0.1 to 10 mm) compared to colloid flotation or applications of DAF for treating
water supplies. In the latter cases, bubbles are at a size of 10 to 100 urn (typically, say 40 urn) and particles
prior to flocculation are 20 urn or less as illustrated in Figure 3b. It is argued below in the modeling section
that flocculated particles need only be tens of micrometres in size, so that in DAF both the particles and
bubbles are tens of micrometres as illustrated in Figure 3b.

Liquid 0)
a)

b) BUBBLES 10 to 100 11m


Typical 40 11m

PART IC LES 0.01 to 10 11m


Desire 20 11m

Pa rticl e • 20 11m
Bubble 0 40 11m

Collisio n Bub ble-Parficle


and Rise
Attachment ::::::>
(j Q
Fig. 3. (a) Illustration of the contact angle for bubble attachment to a solid surface; (b) Size range and scale of
bubbles and particles.
Dissol ved air flotation 5

Derjaguin et al. (1984) present a different view of particle-bubble attachment of colloidal particles by small
bubbles where a finite contact angle need not form, Essentially. their model is a heterocoagulation one
involving the stability of particles and bubbles. Attachment requires reduction in electrical charge
interactions and attraction by London-van der Waals forces as particles are transported to bubble surfaces.
Small bubbles such as those occurring in DAF carry a negative charge (Ekserova and Zakharieva, 1975;
Okada and Akagi, 1987) with a reported IEP (isoelectric point) of 4.5. Small bubbles may adsorb surfactants
or natural organic matter (NOM) and obtain a charge. For waters free of such materials. it is hypothesized
that at the water-bubble interface anions reside closer to the bubble surface than the larger hydrated cations
causing a small negative zeta potential.

No matter which view is taken (the contact angle or the heterogeneous system of small particles and
bubbles). there is universal agreement and experimental evidence that two conditions are necessary for
favorable flotation: (I) charge neutralization of the particles; and (2) production of hydrophobic particles.
These are addressed next.
2.

2.
::- ~
1.
'0

.D
...
::>
Z
~ 1.
...
:>

o.
o.o+---+--+- ~ -t---+---+-----,f--I--+--~
2. (h----,--.,.-- ~-- .- - - , - ..,.-- ---,- - ---,, -- .,.-- -,-- --.,

e 1.
....u
:x: ....>
e,
o.
'" ...."e
~ - 1.

-2.o+---t---+- - +- - -t-- - - + - - - l - -I-- I - - + - --f--'-

o TOlal

.~ o Algal ·Slze
Z ...J
e "&-~--_ ._ ~
""
o
...J
....-
'0 "
.--.- ..-.
~
... - - u -..-.. ., ._._ .._ _{)

-+----+- --; - - .-1-.. -+-I---+Ir--T---+----t!~-l


! !
0 .0 0 .5 1.0 1.5 2.0 25 3 .0 3.5 4.0 4 .5 5.0

PA CL DOSE (mgjL a s Al )

Fig . 4. Effect of coagulant dose on particle charge <EPM data) and DAF clarification (turbidity and particle
number) (Wachusett Reservoir. Boston. USA) Chlorella vulgaris at 10 5 cellsJml added to raw water. pH 5.5.
flocculation time of 5 min. recycle ratio of 8%. Cblorella size 3 to 8.7lJlI1.total size range 2 to 28 lJlII.

Charge neutralization. The data in Figure 4 show the effect of coagulant dose (polyaluminium chloride.
PACI) on the flotation of a green alga (Chiorella vulgaris) at pH 5.5. The PACI was prepared with a OHlAl
ratio of 2.5 so the product was 90 % positively charged polymeric Al species that accomplished particle
destabilization via charge neutralization. Flotation is not successful without the addition of coagulant
Favorable flotation conditions occur at a PACI dose of about 0.5 mg/L as AI. Here. the negative charge on
the Chlorella is reduced (although complete charge neutralization is not necessary) as indicated by the EPM
(electrophoretic mobility) data. and flotation performance is good as demonstrated by the nearly two log
6 J.K.EDZWALD

reduction in algal cells and reduction in turbidity. Overdosing with PACI can occur due to charge
restablization of the particles (positively charged particles and bubbles).

= AIOHj - -(HOH)
pH<ffiP

=AlOH . - - (HOH)
OH
OH I OH pH=ffiP
I I

/ v- /I~ ---/\
Particle Surface =AIO· -- - (HOH)
@IEP
pH>IEP
Fig. 5. Hydrophilic nature of aluminium hydroxyide particles.

~A ... BEFORE OAF


0/ ----o.....~
12
0 AFTER DAF
-' .....~ ~
..... a- -a

O'---'-----'-- -L- .L-----'

... BEFORE DAF


16 o AFTER DAF
~2 12 ",~.
'0......... :& Ja.~.. ~.A
.
i'?6 8
4 20 c e
o ' - - "L -
-1J-

L-
-----0---0__ .r:..=:.=a.J
.L-
-

E
~.~ .5
w-<e.
j 0

-.5'--~'------J~---!:_---_!:_---'
5 6

pH
Fig. 6. DAF turbidity and charge data (EPM) for freshly precipitated aluminium hydroxide in pure water.

Hydrophobic uarticles. Bubble adhesion or attachment to particles requires hydrophobic particle surfaces
(Gochin and Solari, 1983) or hydrophobic spots on particles. For many particles, hydrophobicity is increased
by simply reducing the negative charge. Other particles such as freshly precipitated or amorphous
Al(OHh(am} have polar surface groups that bond water making them hydrophilic. This hydrophilic effect
may be reduced by charge neutralization, but aluminium hydroxide contains a polymolecular layer of water
at the particle surface which hinders bubble adhesion. This layer and pH effects on the isoelectric point (IEP,
pH at zero net charge) are illustrated in Figure 5. To demonstrate the hydrophilic effect with freshly
precipitated aluminium hydroxide, flotation experiments as a function of pH were run in ultra-clean water
(Super Q water; prepared by treatment of reverse osmosis, ion exchange, carbon adsorption, and 0.2 11m
filtration), Figure 6 presents the flotation results and EPM data (particle charge), The EPM data shows that
the IEP occurs at about pH 8. There is improvement in flotation at 200C (measured by the change in
Dissol ved air flotation 7

turbidity before and after flotation) when the pH is increased to values near the IEP, although even at the
IEP the residual turbidity following flotation is about 1 to 2 NTU. At 4°C, flotation is poor even at the IEP
demonstrating the hydrophilic nature of aluminum hydroxide. For particle-bubble attachment to occur, water
must be displaced between the particle and bubble as they approach. At cold water temperatures, this
displacement or thinning of the water is retarded in addition to the hydrophilic effect of the bonded water.

Bubbles supplied. Injection of the pressurized recycle water stream into the flotation tank causes air bubbles
to be released. The concentration of supplied air bubbles that results from mixing the recycle water with the
water undergoing treatment affects particle-bubble collisions, and particle separation and removal. It is thus
an important design and operating parameter. There are three fundamental measures of the air supplied: (1)
the mass concentration (Cr); (2) the air bubble volume concentration (cI>))l; and (3) the bubble number
concentration (N b). In practice the recycle ratio is used as a surrogate measure of the air supplied for a
constant saturator pressure. The recycle ratio is defined as

(2)

Qr = the recycle flow


00 = the influent flow.
Henry's Law is used to calculate the equilibrium mass concentration of dissolved air for water leaving the
saturator. Assuming air is a single ideal gas

(3)

Csat = saturated mass concentration of air in saturator recycle water


f = the saturator efficiency
Pt = the total air pressure of the saturator (i.e., gauge pressure plus one atmosphere)
Hair = Henry's Law constant (4.18 kPalmg/L at 20OC).

Air is a mixture of gases consisting primarily of nitrogen and oxygen. Oxygen has a higher solubility than
nitrogen and dissolves more rapidly producing a nitrogen enriched atmosphere in the enclosed saturator. At
equilibrium, the saturator air is approximately 88% nitrogen and 12% oxygen causing a reduction by about
9% in the dissolution of air compared to 100% saturation with atmospheric air. The efficiency factor (f) can
be used to account for both the nitrogen enriched atmosphere problem and mass transfer efficiency
differences between unpacked and packed saturators. Saturator efficiencies may be 60 to 70% for unpacked
saturators and about 90% for packed saturators.
___ ge9£1<:.- ~ _
(After OAF or Filtration) l
Saturator
'Q
I r
I

Rapid Mix Flocculation DAF


f.-- PRETREATMENT -----j

.0·.··
..
"
:~ :
~
of! ·····
w ....

..

..
..

<10
....
0

o--o--o---o--a
....

<I
..

"
0

<I "
..

..
"

..
..

<I
..
<I

C
..

IReaction
Zone
I Separation
Zone
Reaction Zone

Fig. 7. Schematic of a OAF plant (bottom shows OAF tank divided into reaction and separation zones).
JWST 3/4-8
8 J. K. EDZWALD

The fundamental parameters of mass, volume, and bubble number concentrations in the flotation tank are
calculated from Eqns. 4 through 6. Eqn. 4 is obtained from a mass balance following air injection and
expresses the mass of air per unit volume of water (Cr) in the reaction zone of the flotation tank (Figure 7).

C sat = the mass concentration in the recycle flow from Eqn. 3


C a = the air concentration remaining in solution at atmospheric pressure
k =flotation tank influent flow saturation factor given by (Ca-C o)' where Co is the mass air concentration of
the influent flow
Psar =the density of air saturated with water vapor (1.19 mglcm 3 @ 20OC).

"r--~-~-~--~-~-~------,

12
c
~

;;
10

/
70% Saturator
Efficiency

.
8 10 '2 ,. ,6
Recycle 7;

70% Saturator
Efficiency

e '0 ,2
Rl1l;yt;le 7.

70% Saturator
Efficiency

~.
o _ _~ _ - . J
,.
L-_~_~ ~ ~

o '0 .2 16

Fig. 8. Bubble mass. volume and numberof concentrationsas a function of recycle for two efficiencies (saturator
guage pressure of 483 kPa, 20'C).
Dissolved air flotation 9

Figure 8 shows the mass (C r ). bubble volume (<%>b)' and bubble number (Nb ) concentrations as a function of
the recycle ratio R, for a con stant saturator pressure of 483 kPa. Table 1 presents a summary of calculations
from Eqns. 4 through 6 for typical recycle ratios of 5 to 15%. It shows mass concentrations of 3.5 to 10 mg/l ,
bubble volume concentrations of 3000 to 8000 ppm, and bubble number concentrations of 105 bubbles/ml or
greater. Higher concentrations are obtained with increasing recycle or increasing saturator efficiency.
Reaction zone modeling presented next will show that <%>b and N b are fundamental parameters relating
particle-bubble collisions.

Table 1. Bubble Mass, Volume and Number Concentrations*

Cr <Db
mg/L ppm

5 3.50 2900 8.75 x 104


8 5.45 4600 1.2 x 105
10 6.68 5600 1.7 x 105
15 9.59 8000 2.4 x 105

*Overall efficiency of 70% for dissolving air and for


the saturator, 20°C .

Flotation M Qdelin~

The flotation tank is divided into two zones as illustrated at the bottom of Figure 7. The purpose of the
reaction zone is to contact and adhere particles to bubbles , and the purpose of the separation or clarification
zone is to provide relatively quiescent conditions for particle-bubble agglomerates to rise to the surface.
Edzwald and co-workers (Edzwald et al., 1990; Malley and Edzwald , 1991) have described a conceptual
model for particle collection by bubbles in the reaction zone. A brief summary is presented bel ow followed
by modeling concepts and principles for the separation zone.

Reaction zone model . The removal efficiency (R) of particle s by a single bubble is expressed as

R = Upb 11T (l 00%) (7 )

pb = the adhesion efficiency


11T = the total single collector (bubble) efficiency.

The single collector efficiency concept is used to describe particle transport to bubble surfaces. It has been
used by others for flotation processes (Flint and Howarth , 1971; Reay and Ratcliff, 1973) and as the
modeling approach for removal of particles by deep bed filtration (Yao et al., 1971). The convective-
diffusion equati on and particle trajectory analyses are used to derive individual particle transport expressions
for the single collector efficiency for Brownian diffusion (11 0 ), interception (111), sedimentation (11s) and
inertia (11IN)

(8)

110 = 0.9 (kT/)l d p db Ub)2/3 (9)

111 = 3/2 (dpld b)2 (10)


10 J. K. EDZWALD

TIs = (p p - Pw) g dp21 08 Jl Vb) (11)

TlIN = (g p w db dp2) I (324 2) (12)

k = Boltzmann's constant
T = absolute temperature
dp = particle diameter
Pp = particle density
Jl and Pw are the water viscosity and density
Vb = bubble rise velocity (fluid motion around bubble relative to the bubble)
g = gravitational constant of acceleration.

Transport by inertia is not significant for particles and bubbles less than 100 um, so TlIN is no longer
considered for DAP. It may be important; however, for dispersed air applications where particles and
bubbles are larger.

For Stokes laminar flow conditions (applicable for bubbles of 100 urn and less) the bubble rise velocity is
calculated from

(13)

where Pb is the bubble density (Pw » Pb)' Substitution into Eqns. 9 and 11 yields

TID = 6.18 (kT/g Pw)213 0/dp)213 0/d b)2 (14)

TIs = [(pp -Pw)/Pwl (dJdb)2 (5)

~
=
.!l
u 10-1
IS
~
=
.S!
..,
10-2 .:
TIn
==e
U
.
Su
eu 10-3
=C
U
eu
'Q 10-4
.S
rI:l
.;
10-3 10-2 10-1 101 Hi'
Particle Diameter (~m)
Fig. 9. Single collector efficiency vs particle size (TID is Brownian diffusion, III is interception, TIs is sedimentation;
for particle density of 1.01 g/crn3, bubble diameter of 40 IJ.III and 25°C).

The effect of particle size on TIT is shown graphically in Figure 9 for a bubble diameter of 40 urn. These
calculations show that TIs is not an important mechanism. They also show that TIT has a minimum value for
Dissolved air flotation II

0.9 urn particles (i.e., particles of about 1 urn) with increasing values for smaller and larger particles. llT is
controlled by Brownian diffusion (l1D) for particle diameters below 1 urn, and is controlled by interception
(11,) for particles larger than 1 urn.

Eqns. 10, 14, and 15 are instructive. They show that llT is dependent on bubble size according to d b·2 for all
cases and the dependence on particle size is d p·213 for Brownian diffusion and d p2 for interception and
settling. Hydrodynamic retardation (tendency to decrease particle collisions as particle moves around
bubble) was not included in the model development. If included, the exponent for d p would be slightly less
than 2 for the interception and sedimentation equations. Reay and Ratcliff (1973) ran experiments to test the
collision efficiency dependence on particle size. They examined the removal of glass microspheres
(diameters of 1 to 20 urn) by collection on small bubbles (20 to 100 urn), which were produced
electrochemically. They found good agreement between theory and experimental results with R depending
on particle size to dp2.05 . These results support neglecting hydrodynamic retardation for these size particles
and bubbles.

The following particle removal rate equation is obtained by extending the removal by a single bubble to a
system containing a bubble number concentration of Nb

(16)

Np = particle number concentration


Ab = projected area of the bubble.

This equation may be expressed in a simple rate form in terms of a rate constant k for flotation as

(17)

where k equals (Upb llT) CAb Db Nb). Eqn. 16 may also be rewritten in terms of the bubble volume
concentration and bubble diameter (db) using Eqn. 6 and setting A b equal to (x d b2) / 4

(18)

It can also be expressed in terms of the particle removal per flotation column or tank depth (dH)

(19)

Discussion. Equations 16-19 apply to batch flotation conditions. Nonetheless, the equations can be used as a
conceptual model showing the importance of parameters affecting flotation performance. The right-hand
side (RHS) of Eqn. 19 identifies key design and operating parameters

(<lpb N p ) llT (<I>b / db)


pre-treatment flotation reaction zone

A summary of these parameters is given in Table 2. pb and N p are pre-treatment parameters which are
changed through coagulation and flocculation. Flocculation reduces N p which reduces the collision rate;
however, this coincides with an increase in particle diameter (d.,). T depends on d p2 yielding higher T values
for particles of lO's of micrometers (see Fig. 9). Within Table t, T is shown under the reaction zone section,
but is strongly affected by pre-treatment (flocculation) effects on particle size. Overall, the particle removal
rate depends on db" which is obtained through an analysis of Eqn. 16. This means that smaller bubbles
improve flotation, and is a primary reason why DAF is such an effective process. As explained above, the
bubble size is pretty much fixed between 10 and 100 m. The designer and operator do not vary this
parameter.
12 J. K. EDZWALD

Table 2. Summary of Conceptual Reaction Zone Model Parameters

Parameter Dependence Comments

PRE- TREATMENT PARAMETERS

lXpb I. Particle-bubble I. Favorable flotation: Requires


(particle-bubble attachment charge interactions reduction in particle charge and
efficiency) hydrophobic particles
2. Hydrophilic nature of
particles 2. Increase lXpb to 1: Optimum
coagulation and pH conditions

Np 1. Raw water quality 1. Concentration and size of


(particle number concentration) particles; and cone. of NOM
2. Coagulant type and
conditions 2. Coagulants may add particles

3. Flocculation time 3. Flocculation may reduce N p


and increase d p

REACTION ZONE - FLOTAnON TANK

1. Particle-bubble collisions from 1. Increase llT: Produce floc size


(total singl~Iollector efficiency) diffusion and interception of lO's of microns

2. Minimum llT for d p of .. 1 urn 2. Short flocculation times

db Controlled by pressure difference 1. Desire microbubbles: Range 10


(bubble diameter) across nozzle and injection flow to 100 urn, median 40 urn

2. llT ex: db-2 ; Rate of collection


of particles ex: db- 1

3. Smaller bubbles: Better


performance

<1>b 1. Saturator pressure 1. Increasing <1>b increases Nb :


(bubble volume concentration) More bubbles for collection of
2. Recycle ratio particles

2. Increase C1>},: More bubble


volume for reducing floc
density

The designer and operator can control the concentration of air released in the reaction zone. Expressed
fundamentally in terms of N b and (J)b and comparing them to the particle number (NP> and particle volume
((J)~) concentrations is useful. Table I shows N b and (J)b values for typical recycle rates. N b is in the range of
10 to 2.4 x 105 bubbles/ml and (J)b is in the range of 3000 to 8000 ppm (i.e .. cubic metre if bubble volumel
106 cubic metres of treated water). Surface water supplies may have an extreme range for N p of 10 3 to 106
Dissolved air flotation 13

particles/mi. but a more realistic range to consider is 103 to 105 particles/ml. Coagulants may precipitate and
increase Np but flocculation will decrease it It is reasonable to assume that in most cases N p in the reaction
zone of the flotation tank (i.e.• following flocculation) will be less than 104 to 105 particles/ml., N b is
therefore one to ten times the particle number concentration. The particle volume concentration for water
supplies should be in the range of 0.5 to 100 ppm with most supplies at 0.5 to 10 ppm (suspended solids
mass concentrations of 20 mgll or less). <!lp can be increased by use of precipitating coagulants, but in most
cases the increase would not exceed tens of ppm. With <!lb of 3000 to 8000 ppm and <!lp of lO's ppm. the
bubble volume is much, much greater than particle volume (<!lb» <!lp>.

Tambo and co-workers (Tambo and Matsui. 1986; Fukushi et aI., 1995) take a different approach and model
bubble-particle collisions as a heterogeneous flocculation process. Both the author's approach using the
single collector collision concept and Tambo's approach are reasonable and valid ways to model particle
removal by bubbles. Each approach makes assumptions and each model needs to be applied properly.
Tambo considers correctly that the bulk fluid flow in the reaction and separation zones is characterized by
turbulent flow. His flocculation rate equations for the disappearance of bubbles and particles are written for
collisions caused by bulk water velocity differences from turbulent mixing. This author uses an entirely
different approach for collection of particles on bubbles i.e.• the single collector collision efficiency model.
Here transport of particles is considered in the vicinity of the bubble surface. The relative motion of the fluid
to the bubble is considered to create streamlines of flow around the bubble. These streamlines of flow
around the rising bubble occur under laminar flow conditions (local Reynolds number (Re)). Both flow
conditions are correct. What is different, are the approaches regarding particle transport. An additional point
regarding the author's model. is that it is intended as a batch conceptual kinetic model that must be
incorporated into a continuous flow system for the particular DAF application.

The flotation rate constant (k) obtained from the author's single collector collision efficiency approach per
Eqns. 16 and 17 is

(20)

The actual difference between the author's and Tambo's approaches are contained in llT A b Vb which is the
bubble-particle collision rate. Tambo 's flocculation modeling approach is based on the following rate
equation

(21)

where r pb is the collision rate and l3(dp' db) is a collision frequency function modeled by Tambo as transport
between particles and bubbles due to differences in bulk water velocities from turbulent mixing. The author's
equivalent term for l3(dp• db) is the product llT A b Vb' where collisions at the bubble surface occur primarily
by interception (boudary condition for differences in velocity between bubble and water).

Separation zone. Following bubble attachment and reduction in particle density, particle-bubble
agglomerates rise to the surface of the flotation tank in the separation or clarification zone. Some basic
equations are presented first to illustrate the role of bubbles in reducing floc density and producing flotation.
Eqn. 22 permits calculation of the particle-bubble agglomerate density (Pp~ assuming spherical bubbles and
floc particles of equivalent spherical diametre (dP>

(22)

Pp = floc density of the initial particle


Bn = number of attached bubbles.

The equivalent spherical diameter (dp~ of the particle-bubble agglomerate can be computed from
14 J. K. EDZWALD

(23)

The rise velocity of the particle-bubble agglomerate (vp.,) assuming Stokes drag conditions is calculated by

(24)

Table 3 summarizes calculations for bubbles of 40 m and floc particles initially varying from 10 to 500 m.
These calculations show

one bubble (db) easily reduces the floc density of particles 100 m or less in size;

two or more bubbles are needed to float floc particles of 200 m;

it is better to keep floc particles less than 100 m to obtain high rise velocities;

ten bubbles attached to a 500 m floc particle will not result in flotation (i.e., a positive rise velocity);
at least 24 bubbles are required to float this particle.

Table 3. Particle-bubble Agglomerate Densities, Equivalent Diameters and Rise Velocities (Bubble
Diameter of 40 urn)

Initial dn * One Bubble Attached


Ppb dpb vpb
(urn) (g/cm 3 ) (urn) (mIhr)
10 0.02 40 3.1
20 0.11 42 3.0
50 0.67 57 2.1
100 0.95 102 1.0
200 1.01 200.5 No Flotation
500 1.01 500.1 No Flotation

Two Bubbles Attached Ten Bubbles Attached


Ppb dpb Ypb dpb vpb
Initial dn * (g/cm 3 ) (urn) (mlhr) (urn) (mlhr)
10 0.01 50.5 5.0 0.003 86 14.5
20 0.06 51 4.9 0.01 86.5 14.5
50 0.50 63 3.9 0.17 91.5 13.7
100 0.90 104 2.2 0.62 118 lOA
200 0.99 201 0.3 0.94 205 5.2
500 1.01 500.2 No 1.01 501 No
Flotation Flotation

*Initially particle density at 1.01 g/crn-'. 20°C

Next, the relative number of bubbles to the floc particle concentration is considered. The bubble number
concentration (Nb) is set at 1.2 x 105 bubbles/ml (see Table 1). If the floc particle size is 50 um and the floc
particle number concentration (N p) is 104 particles/ml or less, then there will be at least 12 bubbles for every
particle insuring rapid collection of particles and many more bubbles than necessary to reduce the floc
density and produce a high rise velocity (see Table 3). On the other hand if extensive flocculation occurs
producing 500 urn size floc particles. then there would be insufficient bubbles to float N p at 104 particles/rnl.
Dissolved air flotation 15

The particle number concentration would have to be limited to 5000 particleslmL for flotation to occur,
assuming all bubbles attach uniformly at 24 bubbles per floc particle.

Modeling the clarification step involves use of Stokes law for expressing the particle-bubble agglomerate
rise velocity. Tambo and co-workers (Tambo and Matsui, 1986; Fukushi et al., 1995) describe the rise
velocity in what appears to be a complicated equation; however, the equation simply incorporates the change
in floc size due to particle flocculation and agglomeration with bubbles. It also uses a general expression for
the drag coefficient. Their equation is based, however, on Stokes law and is similar to other expressions
below.

For laminar flow conditions around the particle-bubble agglomerate at it rises (i.e., Re (Reynolds Number)
of 1 or less), then Stokes drag is used and Eqn. 24 is valid. In Table 3, the Re is less than 1 for all cases of
one and two bubble attachment. For 10 bubble attachment, the Re is about 3 for all cases, so use of Eqn. 24
for computing the rise velocities is a reasonable approximation. Schers and van Dijk (1992) present Eqn. 24
for the appropriate Re conditions and the following equation for 1< Re < 50

(25)

where u is the kinematic viscosity. Removal of particle-bubble agglomerates occurs under plug flow
conditions when

(26)

where Yos is the overflow rate of the separation zone (i.e., the hydraulic loading based on the area of the
separation zone), and m is the fraction of dead space. This assumes a critical particle entering at the bottom
of the separation zone. Particle-bubble agglomerates with rise velocities less than Yos may be removed
depending on their position at the entrance to the separation zone. Flotation tanks with baffles directing flow
to the surface improves particle-bubble agglomerate separation with respect to Eqn. 26.

APPLICA nONS

There are two purposes to this section. Both deal with drinking water applications of dissolved air flotation.
First, major design and operating parameters are summarized based on practice in several countries. Second,
the concept of integrated treatment for DAF is presented and illustrated with examples.

Desif:n and Operation: Practice

Table 4 contains design and operating parameters compiled for several countries using several sources.
Before discussing the parameters, a couple of comments about the sources are made. The data for South
Africa were obtained from a recent publication that is a well-documented and an excellent design guide
(Haarhoff and van Vuuren, 1993). This design guide contains information on a survey of practice in South
Africa as well as information for other countries. The information presented in Table 4 is their recommended
design values. Two sources of data were used for OAF plants in the United Kingdom. One source was the
survey results of Longhurst reported in her thesis (Longhurst, 1985) and later published in an abbreviated
form (Longhurst and Graham, 1987). The other source was a comprehensive twelve to eighteen month
evaluation of four OAF plants recently completed by the author (Edzwald et al., I994b).

Flocculation mixing intensity data are not reported in the table, but are discussed here. There are
discrepancies between practice and reports from well-controlled studies. Many plants are built and operated
with two or three stages of flocculation with tapered mixing. The consensus of the work reported in the
literature; however, is that tapered flocculation is not needed and higher Gs (mean velocity gradient as a
measure of mixing intensity) than sedimentation are desired for improved clarification by DAF. Work
performed at the Water Research Centre (UK) in the 1970's showed better OAF performance (i.e., lower
turbidity) with increasing mixing intensity measured as G (Zabel, 1984; Gregory and Zabel, 1990). An
16 J. K. EDZWALD

optimum G of 70 s-I was found for alum as the coagulant independent of water type. Janssens (1991)
reported that optimum G conditions for DAF depend on the type of coagulant used: 70 to 80 s·1 for alum, 70
s-I for ferric chloride, and greater than 30 s-I for polyaluminum chloride. His results are based on pilot plant
studies done in Belgium and Argentina. Research done by 0degaard (1994) on flotation of wastewater
chemically treated with alum supports the above drinking water applications. 0degaard reports an optimum
G of 60 to 80 s-I. Recent research by Edzwald and co-workers (Bunker, et al., 1995; Edzwald, et al., 1994a)
testing a variety of water types (river water of low to moderate turbidity, reservoirs of low turbidity, and
humic water type supplies) found lower turbidity following DAF with increasing mixing intensity for alum
and polyaluminium chlorides of different formulations, The data supports the above work indicating an
optimum G in the range of 55 to about 100 sec-I .

Table 4. Summary of DAF Design and Operating Parameters

Parameter SouthAfricaI Finland2 Netherlands3,4 UK5 UK6 Scandinavia6


Flocculation
Intensity See Text
Time, min 4-15 20-127 8-16 20-29 18-20 28-44
Flotation
Reactionzone
Time, min 1-4 0.9-2.1
Hyd. Load., m/hr 40·100 50-100
SeparationZone
Hyd. Load., mJhr 5-11 2.5-8 9-26
Total Flotation Area
Hyd. Load., m/hr 10-20 5-12" 8.4-10+ 6.7-7+
Time, min 11-18
Recycle, % 6-10 5.6-42 6.5-15 6-10 5-10 10
UnpackedSaturators
Pressure,kPa 400-600 400-550 460-550
Hyd. Load., mJhr 20-60
Time, sec 20-60
Packed Saturators
Pressure,kPa 300-600 400-500
Hyd. Load., mJhr 50-80
Packing Depth. m 0.8-1.2
Saturators" "
Pressure,kPa 300-750 400-800 310-830 480-550

IRecommendedvalues: Haarhoffand van Vuuren (1993,1994).


2Heinanen(1988).
3Wortel(1991).
4Schers and van Dijk (1992).
5Longhurstand Graham (1987).
6Edzwald et al.• 1994.
"Excludes recycle flow.
"Includes recycle flow.
"·Unspecified with respectto unpacked or packedsaturator.
Dissolved air flotation 17

Early use of DAF for drinking water treatment in Scandinavia and then the United Kingdom followed a
practice of long flocculation periods, similar to sedimentation, as shown in Table 4. From a survey of
operating plants in South Africa, Haarhoff and van Vuuren (1993, 1995) found a large range of flocculation
times from 5 to 120 min; however, in the design guide they recommend times of 4 to 15 min (Table 4). The
flocculation times for The Netherlands are also much shorter than what has been done in Scandinavia and
the UK. Janssens (1991) found in pilot plant studies of treating a water supply for Antwerp (Belgium) that
flocculation times of 5-6 min gave good flotation performance. A full-scale plant was built was built and is
in operation with this short flocculation time. The short flocculation times are in agreement with the reaction
zone modeling of the author presented under Principles. Additional discussion is presented below under
Integrated Treatment.

The hydraulic loading or overflow rate is often given as a design parameter for DAF. It and water velocities
at key locations (such as the entrance to the separation zone) are important factors affecting clarification. It
is not always clear whether the hydraulic loading is given for the separation zone or the entire tank. Table 4
shows hydraulic loadings for the reaction zone, separation zone, and the entire tank. Note the high values
and short detention times for the reaction zone indicating that the kinetics of bubble collision and attachment
is fast. Detention times of DAF tanks are small compared to sedimentation tanks.

Haarhoff and Rykaart (1995) address the design of packed saturators. Recommendations on design values
for packed and unpacked saturators are presented in the design guide of Haarhoff and van Vuuren (1993)
and summarized in Table 4. Saturator pressures in Table 4 using data for all the countries may vary from
300 to 800 kPa, but most saturators are designed in the range of 400 to 550 kPa (60 to 80 psig or 4 to 5.4
atm). A unique property of DAF is the microbubbles (10 to 100 11m) formed in the process. To ensure
formation of these microbubbles, the saturator pressure should not be less than the values shown in Table 4.
Pressures in the range of 400 to 600 kPa are recommended. The main way to control and vary the amount of
air supplied to the DAF tank, and thus the concentrations of bubbles, is through the recycle rate defined as
the recycle flow divided by the plant flow or Q/Q o (see Fig. 7). In practice, the data in Table 4 shows that
this varies from 5 to over 40% with most plants at 6 to 10%. Edzwald et al. (1992) examined air
requirements for floating three different water types including a fulvic acid water at DOC concentrations of
2 to 15 mg/l (colour of 20 to 110 Pt-Co units), clay turbidity waters with clay solids at 20 to 100 mg/I, and
algal type waters with algae at cell concentrations of 20,000 to 500,000 cells/ml. In these experiments the
saturator pressure was 483 kPa (70 psig) and the efficiency of air dissolution was 70%. Recycle ratios of 4.5
to 8%, depending on raw water type and the concentration of the contaminant, were required to produce
DAF turbidities if 1 NTU or less. A recycle ratio of 8% (bubble volume concentration (cI>~ of 4600 ppm and
bubble number concentration (Nb) of 1.2 x 105 bubbles/mL) was sufficient to treat successfully all water
quality types and concentrations of contaminants.

Intelirated Treatment

The concept of integrating DAF into an overall treatment plant considers that what is done ahead of flotation
with coagulant addition and flocculation (i.e., pretreatment) affects flotation performance, and in tum,
flotation performance affects filter design and performance. There is a misconception that the role of DAF in
water treatment is limited to supplies containing algae or natural color. By integrating DAF into the overall
plant, it has wide application in treating surface water supplies.

Pretreatment. Particle-bubble interactions and the role of coagulation were discussed above under
Principles. In summary, favorable conditions for bubble attachment to particles requires coagulation
conditions that reduce particle charge and produce hydrophobic particles. Coagulant dosages and pH
conditions that satisfy these criteria depend on coagulant type and raw water quality characteristics,
particularly, particle concentration, hardness, and the concentration and type of NOM. For many water
supplies that have a sufficient concentration of NOM that is humic in nature, hydrophobic particles are
produced with metal coagulants and there is no need for the addition of organic polymers (flotation aids) to
increase the hydrophobicity of floc particles. For this case, the coagulation requirements, or more
fundamentally the particle destabilization requirements, for DAF are similar to other downstream particle
18 J.K.EDZWALD

separation processes such as sedimentation or direct filtration , and therefore coagulation conditions are
similar. An important difference is that flocculation kinetics is important in sedimentation plants often
requiring increasing the coagulant dose for sweep floc conditions, which is not necessary for OAF.

2 .'
.z Tim e Polymer Added Aft er Alu m
..s. 2 .0
o 2 m in
?;- I.. • 5 m in
'U
~ V l0m in
:e
:>
1.2
l-
0 .8
"0
c:
tz 0.' ,
0 .0
200
':J
<,

3'" 160

..,+ 120
:;;'

:0 "
:>
80
(;
Vl 40
(;
c
c;: 0
0 5 '0 15
Flocculoti on Time Afte r Pol yme r Add it ion (min )

Fig. 10. Effect of adding a flotation aid (ani onic polymer) following alum at vari o us times in treating a low DOC
water (Wachusett Reservoir, Boston , USA) under cold water conditions of 4°C (alum at I mg/l as Al, polymer at
0.001 mg/l, pH 6.3).

It was shown above with Figure 6 that aluminium hydroxide particles in clean waters are hydrophilic and
float poorly. For low turbidity waters that are low in NOM, and the composition of the NOM is such that it
is comprised of mostly low molecular weight and hydrophilic organics, then a flotation aid may be
necessary. This is illustrated with data from OAF studies of the Wachusett Reservoir (Boston). This supply
is a high quality water low in DOC and aquatic humic matter. flotation with alum is good under summer
water temperatures , but higher residual turbidities occur at low water temperatures, 4-6~ . This problem is
solved by the addition of a high molecular weight, low charge density anionic polymer following alum as
shown in Figure 10. Good flotation is achieved by adding the polymer at least 2 min after the alum. Good
flotation performance is achieved for a total mixing time of 7 min, 5 min of alum reaction followed by only
2 min of flocculation. The role of the polymer is to adsorb onto the aluminium hydroxide particles and create
hydrophobic spots. It may also have sufficient size to extend from the aluminium hydroxide surface into
solution for distances greater than the adsorbed water layer and attach to bubbles. OAF treatment of clean
waters can also be accomplished by using coagulants that do not produce hydrophilic particles. Many of the
PACI coagulants may be used or ferric salts may produce less hydrophilic particles than alum.

The practice of OAF evolved with flocculation tank detention times of 20 to 45 min, as summarized above,
and is thus quite similar to sedimentation plants. The function of these two clarification processes is
different. In sedimentation, large particles are needed to overcome the small density difference between floc
particles and water so that gravity will overcome fluid drag and buoyancy and cause settling. Large floc are
not needed for flotation since we need to attach bubbles to particles and reduce particle-bubble agglomerate
density to less than water as shown above in Table 3. Also, detachment forces are greater for larger floc. It is
better to keep floc particles less than 100 m to obtain high rise velocities. Theory presented above regarding
particle-bubble collisions suggests that floc particles of 10 to 30 m (pinpoint floc) should be prepared for
flotation. Laboratory and pilot studies of the author (Edzwald and Wingler, 1990; Edzwald et al., 1992;
Bunker et al., 1995) have confirmed that small floc particles are effectively removed by flotation and
flocculation times as low as 5 min are feasible.
Dissolved air flotation 19

Data on particle size distributions through a OAF plant are scarce. Some data are summarized next from an
earlier paper (Edzwald et al., 1992). Pilot experiments were carried out with a 0.17 m 3/min (45 gpm) flow
using a two stage flocculation system (gate flocculators) with a mean detention time of 16 min. Bypassing
one stage permitted study with a mean flocculation time of 8 min. Lake Whitney (New Haven, CT, USA)
was the water supply. It is a supply of low to moderate turbidity, color, and DOC with seasonal algae
problems. During the pilot runs, the raw water quality data were: turbidity 4.7 to 5.6 NTU, UV 0.16 to 0.19
crrr l, and DOC of 4.6 mglI. At the time Asterionella (diatom and filter clogging algae) counts in Lake
Whitney were in excess of 1000/ml. Ferric chloride was used at a dosage of 23 to 27 mglI. Other pilot plant
operating information follows: air eductor saturator system with pressure at 480 to 550 kPa (70 to 80 psig)
recycle rate at 10%, OAF hydraulic loading of 12.2 m1hr (5 gpmlft2 ), and the filtration rate at 12.2 mlhr with
dual media filters containing 0.5 m of anthracite (effective size (ES) of 1.1 mm) over 0.25 m of sand (ES of
0.45 mm). Particle size measurements were made using a system based on light blockage. No particles
greater than 100 m were observed in the raw water or at other locations in the pilot plant. Sampling locations
included: raw water, after flocculation (termed "flocculated"), after flotation ("OAF effluent"), and after
filtration ("filter effluent").

1200
o Raw
D Flocculated
1000 .. DAF Effluent
• Filter Effluent
'3 - 16 min floc
s
<,
800
- - 6 min floc
~
..
CD
(J
600

~e, 400

200

a
3.0

2.5

'8p., 2.0

.8
..!::
1.5
;:I
"0
> 1.0
CD

~...
Cl 0.5
e,

0.0
1 10
Particle Dlameter{}Lm)
Fig. II. Particle number and volume distributions at four locations in a DAF pilot plant for flocculation times of 8
and 16 min (Lake Whitney, New Haven, CT, USA).

Particle number and volume distributions are presented in Figure II. The particle number distribution data
shows that the raw water contains a large number of particles less than 10 11m. The peak at 6.25 11m is
indicative of the presence of algae in the reservoir. Following flocculation, there is a decrease in the number
of smaller particles and an increase in the number of particles greater than 10 11m. This increase is seen more
clearly by examining the particle volume distribution plots which show peaks at 35 11m. Surprisingly, there
is little difference in the particle size distributions following flocculation for the two times, more on this
20 J. K. EDZWALD

below. The turbidity following flotation was 004 to 0.9 NTU for both flocculation times. Turbidities after
filtration were less than 0.1 NTU for both flocculation times and there was no difference in the filter head
loss gradients. Total particle counts across the plant for the two flocculation times were (8 min flocculation
data listed first, then 16 min data): raw water at 4000 and 4000 particles/ml , after flocculation at 2000 and
2000 particles/ml, after flotation at 680 and 280 particles/ml, and after filtration at 30 and 30 particles/mI.
Based on the raw water particle counts , OAF provided a particle reduction of 80 to 90% and the combination
of OAF and filtration provided a particle reduction of over 99.9%.

Clearly, large floc particles are not produced and are not needed for DAF. Other data indicated that the floc
particles for this pilot plant had slight additional particle growth between the flocculation tank and at the
point of recycle injection. Average particle sizes (volume average basis) at this location indicated sizes of 20
11m and 25 11m for the 8 min and 16 min flocculation tank mean times.

The author recommends for drinking water applications that flocculation tanks in dissolved air flotation
plants be designed to produce strong floc with particle size distributions of 10 to 30 IlID. This is based on
theory and the results of laboratory and pilot plant studies. Flocculation tanks with two to three stages (to
approximate plug flow) and total detention time of 5 to IO min at average flow rates (say, 3 to 6 min at
maximum flows) are feasible and recommended. Pilot plant studies found that short flocculation times were
feasible leading to construction of a full-scale DAF plant in Antwerp (Belgium) with a detention time less
than 10 min (Janssens 1991, 1992) supporting this author's recommendation. This recommendation does not
suggest that DAF is not effective in removing larger floc particles. Clearly, larger particles or floc exist in
mining, municipal wastewater, and industrial wastewater applications. It does suggest, that large floc need
not be produced in drinking water applications. While large floc ( > 100 m) may be produced in water
treatment such as reported in the work of Fukushi et al. (1995), it is not necessary to produce these large floc
particles for removal by DAF.

It may be asked that given the experience in Europe of success with DAF at longer flocculation times, why
decrease the time? Most of the DAF plants in Europe are relatively small which diminishes the need to
optimize flocculation tank design and integration of OAF into the treatment plant. The flocculation tank
design should be different than for sedimentation plants. They should be smaller resulting in space and cost
savings. As OAF is considered for larger cities, it is essential that an integrated design approach be taken to
realize the economic advantages of smaller flocculation tanks e.g., an 8 min flocculation time has been
considered in the preliminary design of a 1.7 x 106 m 3/day (450 MOD) plant for Boston (Johnson et al.,
1994).

Filtration. Filtration should be integrated into the design of a OAF plant. OAF is efficient in removing
particles and can substantially reduce the particle load to the filters compared to direct filtration or
conventional plants using sedimentation. This integration permits designing filters at higher rates yielding
space and cost savings. In the case of Boston, deep bed monomedia filters could be designed at 24 m/hr or at
a rate of 20 m/hr BAC filters (biologically active carbon) for a plant using ozone following flotation
(Johnson et al., 1995).

Another example of integrating filtration with DAF is the use of OAF in retrofitting in-line filtration or
direct filtration plants. This is illustrated from the following study. The South Central Connecticut Regional
Water Authority (SCRWA) has an in-line direct filtration facility with a plant capacity of 0046 m 3/s (lOA
MOD) with a limited filtration rate of 7.3 m/hr (3 gpm/ft2) due to the raw water quality. Increasing the plant
capacity to 0.88 m 3/s (20 MOD) with the existing treatment of in-line direct filtration would require
doubling the number of filters. An alternative is to increase plant capacity by placing DAF ahead of the
filters and increase the filter hydraulic loading. Pilot studies were carried out in 1992. Details of the pilot
plant are contained in the paper by Schmidt et al. (1995). Table 5 summarizes the results of three separate
runs in which the filtration rate in the in-line filtration pilot plant was held constant at 7.3 m/hr, which is the
same rate as the full-scale plant. The DAF pilot runs were done at filtration rates of 7.3, II, and 14.3 m/hr.
The UFRV data in all cases show that the DAF plant has the capability to produce at least twice the volume
of water in a filter run than operation in the in-line filtration mode . The OAF plant may operate at a filtration
Dissolved air flotation 21

rate of 14.3 mIhr meaning the in-line filtration plant can be retrofitted increasing plant capacity to 0.91 m 3/s
(20.8 MGD) without the construction of additional filters.

Table 5. Comparison of Filters Run Times and Unit Filter Run Volumes (UFRV) for In-line Filtration vs
Dissolved Air Flotation"
IN-LINE FILTRATION

Pilt. Rate Run time


mIhr he

7.3 24 175
7.3 22 160
7.3 23 170

DISSOLVED AIR FLOTATION

Pilt. Rate Runtime


mIhr he

7.3 67 500
11.0 33 360
14.6 24 340

*Pilot plant experiments at South Central Connecticut Regional


Water Authority, Lakes Glen and Watrous.
"Based on terminal head loss of 2.44 m.

RESEARCH NEEDS

The following lists of research needs are divided into Fundamentals and Applications. They are limited to
the author's knowledge and experience, and are thus a starting point. Suggestions are welcome from
researchers and practitioners.

Fundamentals

Bubble-particle attachment. Simple and direct methods are needed to measure the hydrophobicity of small
particles and freshly formed floc particles.

Reaction zone modeling. Present models need to be examined. Alternative models should be considered.
Testing of the models in the laboratory for verification against theory is required.

Separation-zone modeling. Likewise, present models need to be examined and refined. Alternative models
should be considered. Testing of the models is needed at the pilot and field scales.

Flotation tank. A fundamental analysis of different flotation tank geometries is needed to optimize bubble-
particle contacts and separation. These studies may include contacting particles with bubbles in a separate
chamber, pipe, or tank.

Saturators and recycle injection systems. Theories of bubble formation and attachment to particles should be
combined with the design of these systems.
22 J. K. EDZWALD

Applications

Pretreatment by flocculation. Flocculation tank mixing intensities and detention times should be optimized
with respect to different applications. It must be realized that flocculation objectives for drinking water vs.
municipal wastewater secondary plant effluent vs. stormwater clarification etc. are different, and thus
flocculation tank design should be different.

Stormwater treatment. OAF treatment of these dilute wastes should be attractive compared to sedimentation
technologies. Small-scale and field-scale studies should be done to determine best overall methods,
performance, and costs.

Removal ofCryptosporidium and Giardia cysts from drinking waters. OAF clarification and OAF combined
with deep bed filtration ought to be a better separation method for these cysts than sedimentation type plants.
No data exists in the U.S.A. regarding removal efficiencies of these cysts, yet present and proposed
regulations require their removal by clarification, filtration, and disinfection. Some data collected in the UK
has been reported by Hall et al. (1995).

Technology transfer. There is enormous potential for OAF in North America. Its use has been limited due to
a lack of knowledge of the process by users (municipalities, industries), designers (consultants), and
government regulatory agencies. Information needs to be made available on the performance, design, and
costs of OAF. A design guide tailored to North America, similar to the South African guide (Haarhoof and
van Vuuren, 1993) is needed.

ACKNOWLEDGEMENT

This paper is dedicated to Stacey J. Edzwald. Contributions of the following former students are appreciated:
J. P. Malley, Jr., A. S. Paralkar, K. M. Boudreau, O. Q. Bunker, Jr., S. C. Olson, P. O. Schmidt, C. M.
Tamulonis, J. P. Walsh, B. J. Wingler-Giraldo, C. Yu, J. A. Nilson and C. M. Walsh; and current students:
M. B. Kelley, N. K. Vinod, J. O. Plummer and M. T. Valade. Mr N. K. Vinod was particularly helpful in
preparing the figures.

REFERENCES

Bunker, D. Q., Jr., Edzwald, J. K., Dahlquist, J. and GiUberg, L. (1995). Pretreatment considerations for dissolved air flotation:
water type, coagulants, and flocculation. Wat. Sci. Tech 31(3-4), (this volume).
Clift, R, Grace, J. R. and Weber, M. E. (1978). Bubbles, Drops, and Particles. Academic Press, New York.
De Rijk, S. E., Van der Graaf, J. H. J. M. and Den Blanken, J. G. (1994). Bubble size in flotation thickening, Wat. Res., 28,465-
473.
Derjaguin, B. V., Dukhm, S. S. and Rulyov, N. N. (1984). Kinetic theory of flotation of small particles. In: Surface and Colloid
Science, Vol. 13, E. Matijevic and RJ. Good (Eds.), Plenum Press, New York, pp. 71-113.
Edzwald, J. K., Malley, J. P., Jr. and Yu, C. (1990). A conceptual model for dissolved air flotation in water treatment, Water
Supply, 8, 141-150.
Edzwald, J. K. and Wingler, BJ. (1990). Chemical and physical aspects of dissolved air flotation for the removal of algae. J.
Water SRT-Aqua, 39, 24-35.
Edzwald, J. K., Walsh, J. P., Kaminski, G. S. and Dunn, HJ. (1992). Flocculation and air requirements for dissolved air flotation.
Jour. Amer. Wat. Works Assoc., 84(3), 92-100.
Edzwald, J. K., Olson, S. C. and Tamulonis, C. W. (I 994b). Dissolved Air Flotation: Field Investigations. AWWA Research
Foundation and AWW A, Denver.
Edzwald, J. K., Bunker, D. Q., Jr., Dahlquist, J., Gillberg, L. and Hedberg, T. (1994a). Dissolved air flotation: pretreatment and
. comparison to sedimentation. Proceedings of the 6th International Gothenburg Symposium on Chemical Treatment,
Gothenburg, Sweden.
Ekserova, D. and Zakharieva, M. (1975). Research in Surface Forces, Vol. 4 (B. V. Derjaguin, Ed.), Consultants Bureau, New
York, USA, p. 253.
Flint, L. R. and Howarth, W. J. (1971). The collision efficiency of small particles with spherical air bubbles. Chem Engr. Sci.. 26.
1155-1168.
Fukushi, K., Tambo, N. and Matsui, Y. (1994). A kinetic model for dissolved air flotation in water and wastewater treatment. War.
Sci. Tech., 31(3-4) (this volume).
Dissolved air flotation 23

Gregory, R. and Zabel, T.P. (1990). Sedimentation and flotation. In: Water Quality and Treatment, AWW A. 4th Ed., McGraw
Hill, New York. USA, pp. 367-453.
Gochin, R. J. and Solari. J. (1983). The role of hydrophobicity in dissolved air flotation. Wat. Res., 17,651-657.
Haarhoff, J. and Rykaart, E. M. (1995). Rational design of packed saturators. Wat. Sci. Tech., 31(3-4) (this volume).
Haarhoff, J. and van Vuuren, L. (1993). A South African Design Guide for Dissolved Air Flotation. Report for the Water Research
Commission, WRC Project No. 332, Pretoria.
Haarhoff, J. and van Vuuren, L. (1995). Design parameters for dissolved air flotation in South Africa Wat. Sci. Tech., 31(3-4)
(this volume).
Hall, T., Pressdee, J., Gregory, R. and Murray, K. (1994). Cryptosporidium removal during water treatment using dissolved air
flotation. Wat. Sci. Tech., 31(3-4) (this volume).
Heinanen, J. (1988). Use of dissolved air flotation in potable water treatment in Finland. Aqua Fennica, 18(2), 113-123.
Heinanen, J., Jokela, P. and Peltokangas, J. (1992). Experimental studies on the kinetics of flotation. In: Chemical Water and
Wastewater Treatment II, R. Klute and H.H. Hahn (Eds.), Springer-Verlag, New York. USA, pp. 247-262.
Janssens, J. (1991). The application of dissolved air flotation in drinking water production. in particular for removing algae.
Proceedings ofDVGW Wasserfachliehen Aussprachetagung, Essen. Germany. pp. 229-254.
Janssens, J. (1992. Developments in coagulation, flocculation, and dissolved air flotation. Water Engnr, and Management, 139(1),
26-31.
Johnson, B.A., Gong, B., Bellamy, W., Tran, T. (1995). Pilot plant testing of dissolved air flota tion for treating Boston's low
turbidity surface water supply. Wat. Sci. Tech.. 31(3-4) (this volume).
Kitchener, J. A. (1984). The froth flotation process: Past, present, and future - in brief. In: The Scientific Basis of Flotation, KJ.
Ives (Ed.), NATO ASI Series, Martinas Nijhoff, Boston, USA, pp. 3·51.
Kitchener, J. A. and Gochin, R. J. (1981). The mechanism of dissolved air flotation for potable water: basic analysis and a
proposal. Wat. Res., 15, 585-590.
Longhurst, S. J. (1985). Anassessment of the current design practice and performance of dissolved air flotation clarifiers for water
treatment in Great Britain, M.S. thesis, Imperial College of Science and Technol., London, UK.
Longhurst, S. J. and Graham, N. J. D. (1987). Dissolved air flotation for potable water treatment: a survey of operational units in
Great Britain. Public Health Engineer, 14(6),71-76.
Malley, J. P., Jr. and Edzwald, J. K. (1991). Concepts for dissolved air flotation treatment of drinking waters. J. Water SRT-Aqua,
40,7-17.
0degaard, H. (1994). Optimization of flocculation in chemical wastewater treatment. Wat. Sci. Tech., 31(3-4) (this volume).
Okada, K. and Akagi, Y. (1987). Method and appartus to measure the ~-potential of bubbles. Jour. Chem. Engr. Japan. 20, 11-15.
Reay, D. and Ratcliff, G. A. (1973). Removal of fine particles from water by dispersed air flotation: effects of bubble size and
particle size on collision efficiency. Canad. Jour. Chern. Engr.• 51.178-185.
Rykaart, E. M. and Haarhoof, J. (1994). Behaviour of air injection nozzles in dissolved air flotation. Wat. Sci. Tech.. 31(3-4) (this
volume).
Schers, G. J. and van Dijk, J. C. (1992). Dissolved air flotation: theory and practice. In: Chemical Water and Wastewater
Treatment II, R. Klute and H.H. Hahn (Eds.), Springer.Verlag, New York, USA. pp. 223-246.
Schmidt, P. D., Tobiason, J. E., Edzwald, J. K. and Dunn. H. (1995). DAF treatment of a reservoir water supply: comparison with
in-line direct filtration and control of organic matter. Wat. SCI. Tech.. 31(3-4) (this volume).
Sulman, H. L., Picard, H. F. K. and Ballot, J. (1905). U.S. Patent 835,439,29 May 1905.
Takahashi, T., Miyahara, T. and Mochizuki. H. (1979). Fundamental study of bubble formation in dissolved air pressure flotation.
Jour. Chem. Engr. Japan, 12, 275-280.
Tambo, N. and Matsui, Y. (1986). A kinetic study of dissolved air flotation. World Congress of Chem. Engr., Tokyo, Japan. pp.
200-203.
Wortel, N. C. (1991). Flotation in the Netherlands. Research Paper, KIW A, Nieuwegein.
Yao, K. M., Habibian, M. T. and OMelia, C. R. (1971). Water and waste water filtration: concepts and applications. Environ. Sci.
Technol.. 5, 1105-1112.
Zabel, T. (1984). Flotation in water treatment. In: The Scientific Basis of Flotation, KJ. Ives (Ed.), NATO ASI Series, Martinas
Nijhoff, Boston. USA, pp. 349-377.

JWST 3/4-C

Das könnte Ihnen auch gefallen