You are on page 1of 8

Full-Scale Evaluation of Aerated Grit Chambers

Author(s): Liliana Morales and Debra Reinhart


Source: Journal (Water Pollution Control Federation), Vol. 56, No. 4 (Apr., 1984), pp. 337-343
Published by: Water Environment Federation
Stable URL: http://www.jstor.org/stable/25042244 .
Accessed: 08/04/2014 15:00
Your use of the JSTOR archive indicates your acceptance of the Terms & Conditions of Use, available at .
http://www.jstor.org/page/info/about/policies/terms.jsp
.
JSTOR is a not-for-profit service that helps scholars, researchers, and students discover, use, and build upon a wide range of
content in a trusted digital archive. We use information technology and tools to increase productivity and facilitate new forms
of scholarship. For more information about JSTOR, please contact support@jstor.org.
.
Water Environment Federation is collaborating with JSTOR to digitize, preserve and extend access to Journal
(Water Pollution Control Federation).
http://www.jstor.org
This content downloaded from 132.216.227.245 on Tue, 8 Apr 2014 15:00:34 PM
All use subject to JSTOR Terms and Conditions
il
Full-scale evaluation of aerated
grit
chambers
Liliana
Morales,
Debra Reinhart
The
importance
of
grit
removal in wastewater treatment has
long
been
recognized by plant operators
and
designers.
Inad
equate grit
removal can result in
unnecessary
abrasion and wear
of mechanical
equipment, deposition
of
grit
in
pipes
or
channels,
and accumulation of
grit
in anaerobic
digesters
and aeration
basins. Poor
grit
chamber
operation
also creates nuisance con
ditions as a result of the removal of
grit
with
high organic
content.
Present
design
criteria for aerated
grit
chambers call for a
95% removal of 65-mesh
grit (particles
with diameter
equal
to
0.210
mm).
Because it is difficult to measure
grit removal,
this
information is of little value to treatment
plant operators
in
evaluating grit
chamber
performance.
This
investigation
eval
uated the
performance
and
design
of five full-scale aerated
grit
chambers in Atlanta. Based on the
results,
the
design
which
produced
the best
performance
was identified. The effect of
operational parameters
and
physical appurtenances
on
grit
chamber
performance
was also evaluated.
BACKGROUND
In an aerated
grit
chamber air creates a
spiral
flow
pattern
through
the chamber. The
spiral velocity
is controlled
by
the
quantity
of air
supplied.
This
spiral
or roll
velocity
controls the
quantity
and
quality
of
grit
removed. Because
grit
removal is
controlled
by
roll
velocity
instead of
flow-through velocity, grit
removal
efficiency
can be maintained over the entire
design
flow
range.1
Most aerated
grit
chambers are
designed
to remove
particles
0.2 mm or
larger
with
specific gravities
of
approximately
2.65.
Adequate
detention times for the removal of these
particles
range
between 2 and 5 minutes.2"4 However,
if removal of smaller
particles
is
desired,
or if the tank is also for
preaeration, longer
detention times
may
be
required.3
The effective removal of
particles
in aerated
grit
chambers
depends
on
adequate hydraulic
control to minimize short-cir
cuiting.5 Short-circuiting
can be controlled
by appropriate
tank
geometry
and
by
the relative
positions
of the inlet
port
and
outlet weirs.6 The
placement
of
longitudinal
or transverse baffles
with
respect
to flow
may
also be a factor in
preventing
short
circuiting.
It has been
suggested
that
long
narrow aerated
grit
chambers
provide
better removal than
square ones.1 However,
no
quan
titative evidence is available to
support
this contention.
The
quality
of
grit
removed is also affected
by
the
type
of
equipment
used. Some of the most
commonly
used devices are
the air lift
pump,
tubular
conveyor,
water
jet pump,
chain and
bucket,
screw
conveyor
and
grit pump.6
In this
study,
the effectiveness of
grit
removal
equipment,
tank
geometry, baffling
and air flow rates was
investigated.
METHODOLOGY
A detailed evaluation of five
grit
chamber
designs
in
operation
at wastewater treatment facilities in Atlanta was conducted.
Each
facility
has one or more aerated
grit
chamber
units,
each
with
unique configuration
and
equipment.
An initial
survey
of
each
grit
chamber
was made to obtain
engineering drawings,
modifications to the
original design, operator experience
with
the units and
any
evidence from
throughout
the
plant
of poor
grit
chamber
performance.
Each tank was drained to observe
areas of
grit deposition
which indicated areas of low or no
velocity
and thus
provided
a
key
to the effectiveness of the air source
to maintain a uniform roll of wastewater
throughout
the tank.
In
addition, velocity
measurements were made
throughout
the
tanks
using
a
portable
water current meter.
Tracer studies were conducted at each
plant
to evaluate tank
hydraulics.
A fluorescent
dye
was introduced at a
point
of com
plete
mix ahead of the
grit
chamber inlet. A submersible pump
in the effluent channel
continuously
fed a
sample
to a fluorom
eter.
Dye
fluorescence was recorded to
produce
a curve of flu
orescence versus time. The
average
detention time for each tank
was obtained from the curve. Flow
pattern
of the
dye through
the
grit
chambers was observed for evidence of
short-circuiting
and dead zones.
The
shape
of the
grit
chamber is not the
only key
to
good design;
diffuser
placement,
air source,
and
adequate baffling
all affect
performance.
Average
values for the
quantity
and
quality
of the
grit
removed
at each
plant
were obtained from
plant
records. These values
were
compared
to the results of
samples
collected
during
this
study.
The last
step
in the evaluation of the
grit
chamber
performance
was to measure the
percent
removal of
grit particles equal
to
or
larger
than 65-mesh.
Sampling
of
grit
chamber influent and
effluent was not feasible because of stratification of
grit
with
depth
in the influent and effluent channels,
and the
extremely
large quantity
of
sample
needed. Instead a more convenient
method was
developed
to obtain similar results.
It was assumed that all
grit equal
to or
larger
than 65-mesh
coming
into the
plant
would be removed in the
grit chamber,
or would settle in the
primary
clarifiers.
Twenty-four
hour com
posited grit samples
were collected from the
grit hopper
and
sludge
from the
primary
clarifiers. Residue
analyses
were
per
formed on these
samples (volatile
and fixed matter in nonfiltrable
residue and in solid and semi-solid
samples) by
method 209G
April
1984
337
This content downloaded from 132.216.227.245 on Tue, 8 Apr 2014 15:00:34 PM
All use subject to JSTOR Terms and Conditions
Morales & Reinhart
LONGITUDINAL SECTION
1111
PERFORATEp
PIPE
=a
CROSS-SECTION
INLET
Figure
1?R. M.
Clayton grit
chamber.
in "Standard Methods." The
particle
size distribution of the
sample
inerts was obtained
by sieving
each
ignited sample
through
selected U. S. Standard Sieves
using
a Standard Me
chanical Sieve Shaker. The
quantity
of
grit
and raw
sludge
re
moved
during
the
sampling period
and the volume of wastewater
treated were also recorded.
Equations
1
through
5 describe the
procedure
for calculation
of the
percent
removal of inerts
equal
to or
larger
than 65-mesh
across the
grit
chamber.
X
=
Q?TSg)(\
-
TVSg) XVxXSg (1)
where:
X
=
dry weight
of inerts removed over a 24-hour
period,
(kg/d),
Qg
=
wet volume of
grit
removed over the 24-hour
period,
m3/d,
TSg
=
total solids
present
in 24-hour
composite grit sample,
fraction,
TVSg
=
total volatile solids
present
in 24-hour
composite grit
sample, fraction,
Vi
=
specific weight
of
water,
1000
kg/m3,
and
Sg
=
specific gravity
of
grit inerts,
2.5.
Y=
Q5(TSs)(l
-
TVSS) XVXXSS (2)
where:
Y
=
dry weight
of inerts removed
by
the
primary
clarifier
over a 24 hour
period, kg/d,
Qs
=
wet volume of
primary sludge
removed over the 24
hour
period, m3/d,
TSS
=
total solids
present
in 24-hour
composite primary
sludge sample, fraction,
TVSS
=
total volatile solids
present
in 24-hour
composite pri
mary sludge sample, fraction,
and
Figure 2?Utoy
Creek
grit
chamber.
Ss
=
specific gravity
of
primary sludge inerts,
1.4.
A
=
Pa?X)
+
Pas(Y) (3)
where:
A
=
dry weight
of inerts
larger
than size "a"
present
in
the
plant influent, kg/d,
Pag
=
fraction of 24-hour
composite grit sample
corre
sponding
to size "a"
(obtained
from sieve
analysis),
and
Pas
=
fraction of 24-hour
composite primary sludge sample
corresponding
to size "a"
(obtained
from sieve anal
ysis).
PRa
=
[PaJLX)/A]
X 100
(4)
where:
PRa
=
percent
removal of influent inerts of size "a" across
the
grit
chamber.
The
percent
removal of inerts
equal
to or
larger
than 65-mesh
across the
grit
chamber was calculated as:
PR
= (p**+J>i?+V<*
+
P?)X
x ! 00
A + B + C + D
(5)
where a, b, c, and d
represent
the different size fractions of
grit
inerts. The number of terms in
Equation
5
depends
on the
number of sieves used in the
analyses.
DESCRIPTION OF GRIT CHAMBERS
Schematic
diagrams
of each
grit
chamber
design
evaluated
are
presented
in
Figures
1
through
5. The
grit
chambers are
referred to
by
the
plant
at which
they
are located. The most
important
features of each
design
are
presented
in Table 1.
Figure
3?Intrenchment Creek
grit
chamber.
338 Journal
WPCF,
Volume
56,
Number 4
This content downloaded from 132.216.227.245 on Tue, 8 Apr 2014 15:00:34 PM
All use subject to JSTOR Terms and Conditions
_Process
Design
Figure
4?South River
grit
chamber.
Three different tank
geometries
and air flow
patterns
were
studied. The R. M.
Clayton grit
chamber
(Figure 1)
has a
long
narrow
configuration,
unlike the other four tanks evaluated.
The
Utoy
Creek
(Figure 2)
and Flint River
(Figure 5) grit
cham
bers have a
square configuration
with air fed at the tank center.
However relative
inlet/outlet positioning
and
grit
removal
equipment
differ. The Intrenchment Creek
(Figure 3)
and South
River
(Figure 4) grit
chambers also have similar tank
geometry
but different air
positioning, inlet/outlet arrangement,
and
grit
removal
equipment.
As a result of the similarities and differences
among
the
grit
chambers
evaluated,
it was
possible
to examine
the effectiveness of
particular aspects
of each
design
such as
type
of
grit
removal
equipment, baffling
for
hydraulic control,
type
of aeration
device,
and
grit washing.
Adequate equipment
to control air flow rate was not available
at
any
of the facilities evaluated so air flow rates could not be
optimized.
In
plants
with
multiple
units it could not be deter
mined if the air was
evenly
distributed
among
the tanks.
RESULTS AND DISCUSSION
Analysis
of tracer studies. The
output
tracer
response
curves
are
presented
in
Figures
6
through
10. A
summary
of the
pa
rameters
developed
from the tracer studies is
presented
in Table
2. Theoretical calculations of detention times and overflow rates
were made
using
the actual flow rate at the time of the
study
AIR
?
SUPPLY
Figure
5?Flint River
grit
chamber.
and the tank volume. Actual detention times were obtained
by
measuring
the center area of the fluorescence-time curves. The
velocities
reported
in Table 2
represent
the
range
of velocities
measured in a horizontal
plane
and in the direction of wastewater
roll.
The R. M.
Clayton grit
chamber has the
only long-narrow
configuration among
those evaluated. Under ideal
hydraulic
conditions,
flow
through
this
type
of reactor should be
plug
flow,
that
is,
the fluid
particles pass through
the tank and are
discharged
in the same
sequence
in which
they
enter. The
par
ticles remain in the tank for a time
equal
to the theoretical
detention time.
Ideally,
if a tracer
slug
was
injected
into such
a tank to
produce
a
given
influent
concentration,
the trace would
appear
in the effluent at that same concentration in a time
lapse
equal
to the theoretical detention time. The tracer curve would
be
represented by
a vertical
response
located at a time
equal
to
the detention time and
approaching
zero width.
The R. M.
Clayton
tracer
response
curve illustrated in
Figure
6 indicates that some
longitudinal dispersion
of the
dye
occurred
and resulted in a deviation from the ideal
plug-flow pattern
described above. Actual detention time based on centroid of
the curve was 4.6
minutes,
which is
higher
than the theoretical
value of 3.7 minutes. This could have been the result of the
Table
1?Summary
of
grit
chambers
design
features.
Feature R.M.
Clayton Utoy
Creek Intrenchment Creek South River Flint River
Tank dimensions,
m
Width
Length
Side water
depth
Inlet baffle
Effluent baffle
Aeration device
Air header
placement
Grit removal
equipment
Grit elevator
type
Grit
washing provided
3.7
27.2
4.0
no
no
perforated pipe
lateral, perpendicular
to flow
chain and bucket
screw
conveyor
no
7.6
7.6
3.4
no
yes
discfusers
center
grit pump
screw
conveyor
yes
5.5
6.0
2.3
no
yes
discfusers
lateral, parallel
to flow
tubular
conveyor
tubular
conveyor
no
5.5
4.3
2.9
yes
yes
discfusers
lateral, perpendicular
to flow
chain and bucket
chain and bucket
no
8.5
8.5
3.7
yes
yes
air lift
pump
center
air lift
pump
screw
conveyor
yes
April
1984
339
This content downloaded from 132.216.227.245 on Tue, 8 Apr 2014 15:00:34 PM
All use subject to JSTOR Terms and Conditions
Morales & Reinhart
Table 2?Data collected
during
tracer studies.
Parameter R.M.
Clayton Utoy
Creek Intrenchment Creek South River Flint River
Theoretical
hydraulic
detention time, minutes 3.7 7.4 5.0 2.9 35.0
Actual
hydraulic
detention
time, minutes 4.6 3.8 3.6 5.6 7.8
Overflow rate,
m3/m2d
1 409 782 809 1 455 187
Roll
velocities, m/s
Surface 0.21-0.61 0-0.61 0-0.54 0.24-0.61 0.09-0.67
Bottom 0.03-0.24 0.03-0.24 0-0.40 0.15-0.36 0.03-0.15
presence of
stagnant
zones in the tank that
delayed dye
bleed
to the
effluent, settling
and
resuspension
of
organic particles
that adsorbed
dye,
or the dissolution of
initially
undissolved
dye particles
with time. The
peak
of the curve, which
may
better
represent
actual detention time in this case,
was 3.5
minutes,
indicating
little
short-circuiting.
The
Utoy
Creek
grit
chamber was
designed
as a
complete
mix reactor.
Complete mixing
in such a tank would
ideally
produce
an
output
tracer
response
curve with maximum
dye
concentration
occurring
as soon as the
dye
was introduced. This
concentration would decrease
gradually
in an
exponential
func
tion. As shown in
Figure 7,
this was not the case.
Instead,
the
tracer
response
curve
represents
a condition of
arbitrary flow,
that
is,
a flow
regime
somewhere between
plug
flow and
complete
mix. Uneven
dye dispersion
and dead
spaces
were
observed,
indicating
severe
short-circuiting.
These observations were ver
ified with the actual detention time of 3.8 minutes obtained
from the
curve, which was
significantly
lower than the theoretical
detention time of 7.4 minutes. Short
circuiting
was
caused,
in
part, by surges
in air flow to the
grit
chamber.
Results of tracer studies at the Intrenchment Creek
grit
cham
ber are
presented
in
Figures
8a and
8b,
conducted with and
without
aeration, respectively.
Normal
operating procedure
did
not include aeration because
plant personnel
claimed that no
improvement
in
grit
removal could be found when the aeration
was used. The
dye dispersion pattern
was
basically
the same
during
both studies. Short
circuiting
was observed and verified
by
the curve
peak
and its relative
magnitude
with
respect
to
the rest of the curve. Detention times derived from the tracer
curves were 2.2 and 3.6 minutes for the test with and without
air, respectively.
In both
cases, the actual detention times were
lOOn
5 10 15
TIME,
MIN.
Figure
6?R. M.
Clayton
tracer curve.
significantly
shorter than the theoretical value of 5.0 minutes.
It was clear
during
the tracer studies that surface velocities
through
the tank were increased
by
air
input.
Because of the
relative
inlet/outlet positions,
this air
input encouraged
short
circuiting through
the tank.
The South River
grit
chamber
design
is similar to that at
Intrenchment Creek
except
in the
position
of the air header in
relation to the
inlet,
and the addition of an influent
turning
baffle. However the results obtained from the tracer studies
(Figure 9a)
were
very
different from those obtained at Intrench
ment Creek. The curve
represents
a condition of
arbitrary
flow
where
mixing
seems to be
fairly
uniform
throughout
the tank.
The actual detention time obtained from this curve was 5.6
minutes. This
value,
as for the R. M.
Clayton grit chamber,
was
higher
than the theoretical detention time of 2.9 minutes.
A second tracer
study
was
performed
at the South River
facility,
this time without air
(Figure 9b).
The blowers were
turned off 24 hours before
testing
and
during
tracer tests. With
air,
the
only major
difference between South River and In
trenchment Creek
grit
chambers was the influent baffle. The
results indicated whether a baffle at Intrenchment Creek would
improve performance
or whether the air header
position
needed
to be
changed.
The
output
tracer
response
curve is
presented
in
Figure
9b. The curve has a
shape
similar to that at Intrench
ment Creek with the air on
(Figure 8a).
An actual detention
time of 1.4 minutes was
calculated, suggesting
that the air header
position
was the critical factor.
Moving
this header to create a
wastewater roll
perpendicular
to the
flow-through length
of the
tank should
improve performance. Although
not
crucial,
the
addition of an influent baffle to
encourage
the
spiral
roll should
also
help
reduce
short-circuiting.
-
IOOt a
w
1
<->
1 i
?J I
q: uj
-
\
o o -
\
-I
UJ
-
i
\
O-Ll-.-1-r
5 10 15
TIME,
MIN.
Figure 7?Utoy
Creek tracer curve.
340
Journal
WPCF,
Volume
56,
Number 4
This content downloaded from 132.216.227.245 on Tue, 8 Apr 2014 15:00:34 PM
All use subject to JSTOR Terms and Conditions
Process
Design
100-,
LU
o
2
LU
LU2
o: tu
oo
D^
-1LU
U_CL
m
5 10 15
TIME,
MIN.
100
LU
u
2
LU
or
lu
IJLU
U_CL
HI
5 10 15
TIME,
MIN.
20
Figure
8?Intrenchment Creek tracer curves.
The tracer
response
curve for the Flint River
grit
chamber
is
presented
in
Figure
10. The
shape
of the curve indicates
very
good mixing throughout
the tank. The flow
pattern
differs
sig
nificantly
from that at
Utoy Creek, although
the
designs
are
very
similar. Differences are attributed to an influent baffle at
Flint River which directs the wastewater to the center of the
tank,
and into the air induced
pumping
zone.
However,
even
with the
baffle,
short
circuiting
was
severe,
as indicated
by
an
actual detention time of 7.8 minutes versus a theoretical de
tention time of 35 minutes. Poor
velocity
control in the tank
was also evident when the tank was drained and accumulations
of
grit ranging
from 0.15 to 1.5 m
(6
in to 5
ft)
in
depth
were
found all over the tank bottom.
Percent removal of 65-mesh
grit.
A
summary
of the results
of the
grit
removal studies is
presented
in Table 3. A 95% removal
of 65-mesh
grit
was achieved at three of the five
plants
and is
thus a valid
design
criterion. However it is not a
practical pa
rameter for
performance
evaluation in the field.
Data in Table 3 indicate that the Flint River
grit
chamber
has the best removal
efficiency.
However this
grit
chamber was
operated
at overflow rates much lower than
typical
values of
1 426
m3/m2
d
(35
000
gal/day sq ft). Large
accumulations of
grit
in the tank bottom are evidence of the failure of the Flint
River
design
to establish a desirable flow
pattern throughout
the
tank;
short
circuiting
is the
apparent
result. Under
loading
conditions
comparable
to those at the R. M.
Clayton
and South
River
installations,
the Flint River
grit
removal
efficiency
would
no doubt be much
lower, perhaps approaching
that found at
Utoy
Creek.
The Intrenchment Creek
grit
chamber also removed a
high
percentage
of the
incoming grit
at an intermediate overflow rate
I0O-1
y*
LU
/
\
<->
/ \
2
\
LU
V
LU2
\
(ELU
\
-J?J l
LLO. \
0-u-,-,-^-r
5 10 15 20
TIME,
MIN.
Figure
9?South River tracer curves.
of 809
m3/m2'd (19
860
gal/day-sq ft)
However
significant
short
circuiting
occurred. Grit accumulations
ranged
from 0.30
to 0.76 m
(1
to 2.5
ft)
in
depth,
which indicates
poor hydraulic
control. In order to
compare
these results with those from the
South River and R. M.
Clayton
facilities at
higher
overflow
rates, grit
removal
quantities
were measured at an overflow rate
of 1 572
m3/m2-d (38
580
gal/day-sq ft).
A 62% reduction in
the
quantity
of
grit
removed was the result. This indicates that
at more
typical operating
conditions the
negative
effects of short
circuiting
and
inadequate inlet/outlet arrangements
would be
more
apparent.
As
previously discussed,
the flow
pattern
at the South River
grit
chamber when
operated
without air was
very
similar to that
at Intrenchment Creek. Grit removal
efficiency
at the South
River
grit
chamber with the air off was
82.7%,
a lower removal
than the 94.8%
during
normal
operating
conditions. This also
suggested
that
performance
would decrease
significantly
at In
trenchment Creek under
higher loading
conditions.
Therefore,
Intrenchment Creek was eliminated as an
optimum design.
For
typical
detention times and overflow rates, grit
removal
at the R. M.
Clayton grit
chamber
(long,
narrow
configuration)
seemed to be the best
among
the
grit
chambers
evaluated,
both
in terms of
process efficiency
and
grit quality.
The unit did not
have
any
unusual
operational problems,
and the absence of
grit
deposits
in the tank bottom indicated
good spiral
roll of the
wastewater.
The South River
grit
chamber achieved a removal of 94.8%
during
the
sampling period.
However the
grit quality
was
very
poor
because of
high volatility
and moisture content,
as seen
in Table 3. This resulted in
disposal problems, including
bad
odors, poor
aesthetics around the
grit
collection area, and in
lOO-i
?J
o
z
?J
?J2
C?UJ
ZJUJ
U-?L
5 10 15
TIME,
MIN.
~20
April
1984
341
This content downloaded from 132.216.227.245 on Tue, 8 Apr 2014 15:00:34 PM
All use subject to JSTOR Terms and Conditions
Morales & Reinhart
-
IOOt
^->
LU
/ \
O / \
2
-
/ \
Ui
/ \
%? ' l
\
UJ2
}
\
oruj I
1
oo .
/ I
"DOC
\
U"Q"
/
\
0-1?i-,-,-r-^-,
5 10 15 20
TIME,
MIN.
Figure
10?Flint River tracer curve.
creased
trips
to the landfill. The
poor grit quality
was related
to the
type
and
arrangement
of
grit
removal
equipment
rather
than tank
configuration.
Effect of
operational parameters
on
performance.
Velocities
were measured in all
grit
chambers to determine actual flow
patterns
in the tanks and the effect of
changes
in
velocity
on
performance.
As shown in Table
2,
a wide
range
in
velocity
was found at each
facility,
but there were no
significant
variations
from one
plant
to another. The
velocity changes
did not seem
to affect the overall
performance
of the units. However areas
of low or zero
velocity
caused
grit deposition
in the tanks.
Equalization
of air flow rates
among parallel
tanks was not
possible
because there were no air flow meters at the
grit
cham
bers in
any
of the
plants.
Air flow rates were
manually adjusted
in
response
to
velocity
measurements so that
parallel
tanks had
approximately
the same roll velocities. This
procedure required
significant
effort and could not
easily
be
performed by
the
op
erators.
Unequal
air distribution seemed to result in different
degrees
of removal
among
the various
operating tanks,
which
would in the
long
run cause
grit
accumulation in
subsequent
treatment units.
The overflow rate data
presented
in Table 2 and the
percent
removal data in Table 3 show that overflow rate is not crucial
to
good removal, provided
air header
placement,
air flow
rates,
and
baffling
are
adequate.
Where these factors were
inadequate,
extremely
low overflow
rates,
such as those found at Flint
River,
seemed to
compensate
for the lack of an efficient
design.
Effect of
physical appurtenances
on
performance.
The effective
operation
of an aerated
grit
chamber
depends
on
many factors,
some of which are
very
difficult to evaluate. Good
performance
requires
not
only
efficient
operation,
but also a unit which is
relatively trouble-free, provides enough operational flexibility,
and
yields good quality grit
that can be
easily
handled and
disposed.
Results indicate that factors such as diffuser
placement,
type
of
grit
removal
equipment,
air flow
rate,
and
baffling
were
of
greatest importance
in
achieving
these
goals.
The
position
of the air headers in relation to the direction
of flow
through
the tank was crucial in
establishing
a
proper
flow
pattern.
Air header
placement producing
a
spiral
roll of
the water
perpendicular
to the flow
through
the
length
of the
tank
(R.
M.
Clayton
and South
River)
seemed to minimize
short
circuiting.
Center air feed
(Utoy
Creek and Flint
River)
and air headers
placed
to
produce
a wastewater roll
parallel
to
the direction of flow
through
the tank
(Intrenchment Creek)
promoted short-circuiting.
The use of
inlet, outlet,
and
longitudinal
baffles
produced
better tank
hydraulics.
At the Flint River
grit chamber,
a baffle
at the
submerged
inlet
port
directed the
incoming
water toward
the center of the
tank,
where a
spiral
roll could
effectively
be
induced. At
Utoy Creek,
which had similar tank
geometry
and
air header
placement,
the lack of such a baffle diminished ef
fectiveness of the air source. The lack of an inlet baffle at the
R. M.
Clayton grit
chamber was offset
by
a
high length-to-width
ratio. At the South River
grit chamber,
a baffle at the inlet
provided
the
hydraulic
control
necessary
to
prevent
short-cir
cuiting
across the
length
of the tank. A
longitudinal
baffle
placed
parallel
to air headers at the R. M.
Clayton grit
chamber allowed
better water circulation and reduced
channeling.
Thus
hydraulic
control can be
improved by
the installation of baffles where
necessary.
Water content of removed
grit
is related to the
type
of
grit
removal
equipment
used. Screw and tubular
conveyors (R.
M.
Clayton
and Intrenchment
Creek) discharged
a
very dry grit.
Grit
washing (Utoy
Creek and Flint
River)
did not seem to
guarantee
low
organic
content in the
grit.
In a
long
narrow tank
such as that at the R. M.
Clayton plant,
the extra
length provided
a
washing
effect that
produced grit
of
very
low
organic
content.
At the South River
grit
chamber the use of chain and buckets
to remove and
discharge
the
grit directly
into a
hopper
without
washing
caused excessive water
carry-over
and
grit
with
high
volatility.
CONCLUSIONS
From the observations and data collected in this
investigation,
the
following
conclusions can be made:
A 95% removal of 65-mesh
grit
can be achieved in con
ventional aerated
grit
chambers and is a valid
design
criteria.
Table 3?Grit chamber
performance
data.
Parameter R.M.
Clayton Utoy
Creek Intrenchment Creek South River Flint River
Grit removal,
wet
basis,
cm3/m3
treated wastewater 10.9 6.7 39.1 17.1 34.0
Grit
quality
TS,
% 79 70 65 28 53
TVS,
% 8 22 13 38 26
Percent removal of inerts
equal
or
larger
than 65-mesh 95.4 82.0 97.6 94.8 98.8
342 Journal
WPCF,
Volume
56,
Number 4
This content downloaded from 132.216.227.245 on Tue, 8 Apr 2014 15:00:34 PM
All use subject to JSTOR Terms and Conditions
Process
Design
The removal of 65-mesh
grit
in full-scale
grit
chambers can
be measured
through sampling
of the
grit
removed and
primary
sludge.
However this
methodology
is not
practical
for routine
performance
evaluation.
Tracer studies are an excellent tool for
performance
eval
uation and
troubleshooting
of aerated
grit
chambers because
they easily identify
areas where corrective action is needed to
improve performance.
The
shape
of an aerated
grit
chamber is not
necessarily
the
key
to
good design. Proper
diffuser header
placement,
a
steady
air
source,
and
adequate baffling
for
hydraulic
control are es
sential in
establishing
uniform flow
pattern,
which will
improve
performance.
Overflow rate does not seem to be a critical
parameter
in
the removal of 65-mesh
grit, provided
air header
placement,
air flow
rates,
and
baffling
are
adequate.
Among
the
designs
evaluated and for a
typical
detention
time in the
range
of 3 to 5
minutes,
a
long,
narrow tank
provided
the best unit
process efficiency, grit quality,
ease of
operation,
and
adequate hydraulics.
Such tank
performance
was less de
pendent
on the use of influent and effluent baffles than that of
square grit
chambers.
Square grit
chambers were effective in the removal of 65
mesh
grit only
if the air headers were
positioned
to create a
spiral
roll of the water
perpendicular
to the direction of flow
through
the
tank,
air flow rates were
adequate
to maintain uni
form velocities
throughout
the
tank,
and
adequate hydraulic
control was
provided by
the use of baffles.
Water content and
volatility
of the
grit
were related to the
type
of
grit conveyor
and washer used. Screw and tubular con
veyors discharged
the
dryest grit.
Grit
washing
increased the
water content
significantly.
ACKNOWLEDGMENTS
Credits.
Operating personnel
of the
City
of Atlanta water
pollution
control
plants provided
assistance. W. H. Cross of
Georgia
Institute of
Technology
and D. Sumlin and R.
Bullard,
directors at laboratories of the Bureau of Pollution Control and
Bureau of Water of the
City
of
Atlanta, respectively, provided
some of the
equipment
and materials. Partial results of this
investigation
were
presented
at the 1981
Georgia
Water and
Pollution Control Association Annual Conference.
Authors. At the time of
investigation,
Liliana Morales was
research
engineer
at the Division of Research and
Development,
Bureau of Pollution
Control, City
of Atlanta. Debra Reinhart
is chief of the Division of Research and
Development.
Corre
spondence
should be addressed to Liliana
Morales,
CH2M
HILL,
1941 Roland Clarke
Place, Reston,
VA 22091.
REFERENCES
1. "Wastewater Treatment Plant
Design,
MOP 8." Water Pollut. Con
trol
Fed.,
and Am. Soc. Civil
Eng.,
Lancaster
Press, Inc., Lancaster,
Pa.
(1977).
2. Metcalf and
Eddy, Inc.,
"Wastewater
Engineering: Treatment,
Dis
posal,
Reuse."
(2nd Ed.) McGraw-Hill,
Inc.
(1979).
3.
Neighbors,
J.
B.,
and
Cooper,
T.
W., "Design
and
Operation
Criteria
for Aerated Grit Chambers." Water Sew. Works, 112,
12
(1965).
4. "Procedures for
Evaluating
Performance of Wastewater Treatment
Plants." Environ. Prot.
Agency,
Office of Water
Programs,
Wash
ington,
D. C.
(Contract 68-01-0107).
5.
Finger,
R. E.,
and
Parrick, J., "Optimization
of Grit Removal at a
Wastewater Treatment Plant." / Water Pollut. Control Fed., 52,
8
(1980).
6. Albrecht,
A.
E.,
"Aerated Grit Chamber
Design
and
Operation."
Water Sew. Works, 114,
9
(1967).
April
1984
343
This content downloaded from 132.216.227.245 on Tue, 8 Apr 2014 15:00:34 PM
All use subject to JSTOR Terms and Conditions

You might also like