You are on page 1of 57

HAZARDOUS

WASTE
MANAGEMENT
VOLUME II
HAZARDOUS
WASTE
MANAGEMENT
VOLUME II
Characterization and Treatment
Processes

SUKALYAN SENGUPTA
Hazardous Waste Management, Volume II: Characterization and Treatment
Processes
Copyright © Momentum Press®, LLC, 2018.

All rights reserved. No part of this publication may be reproduced, stored


in a retrieval system, or transmitted in any form or by any means—­
electronic, mechanical, photocopy, recording, or any other except for brief
quotations, not to exceed 250 words, without the prior permission of the
publisher.

First published by Momentum Press®, LLC


222 East 46th Street, New York, NY 10017
www.momentumpress.net

ISBN-13: 978-1-94561-290-9 (print)


ISBN-13: 978-1-94561-291-6 (e-book)

Momentum Press Environmental Engineering Collection

Cover and interior design by S4Carlisle Publishing Services Private Ltd.,


Chennai, India

First edition: 2018

10 9 8 7 6 5 4 3 2 1

Printed in the United States of America


Abstract

Environmental engineers are responsible for cradle-to-grave handling and


management of a hazardous waste. To fulfill this responsibility, a practic­ing
engineer needs to apply their knowledge of federal, state, and local regula-
tions; environmental audits; toxicology; site characterization; and treatment
processes to transform the hazardous waste site to a condition where it can-
not cause adverse effect to human health and the environment.
Volume I of this series covered the regulatory landscape, basic
­environmental chemistry principles, fate and transport of contaminants,
toxicology, and risk assessment. This second volume focuses on treat-
ment technologies that are com­monly applied at hazardous waste sites and
site characterization. It covers physicochemical processes (air stripping,
­adsorption, ion exchange, and reverse osmosis), incineration, stabiliza-
tion and solidification, biological treatment, and land disposal. Numerous
solved examples provide a step-by-step approach to apply these technol-
ogies in real-life situations. The two volumes combined present a clear
roadmap to the reader to integrate these topics in practice.

KEYWORDS

adsorbate; adsorbent; adsorption; biokinetics; Henry’s constant; HTU; in-


cineration; inhibition; ion exchanger; isotherm; landfill; NTU; osmotic
pressure; permeability; selectivity; separation factor; SRT; stabilization
and solidification; stripping factor
Contents

Acknowledgments ix
Chapter 1 Physicochemical Treatment Processes 1
Chapter 2 Incineration 43
Chapter 3 Stabilization and Solidification 55
Chapter 4 Biological Treatment 63
Chapter 5 Land Disposal 91
Chapter 6 Site Characterization 99
About the Author 105
Index 107
Acknowledgments

The author expresses his gratitude to Professor Francis Hopcroft, editor,


whose constant support, editorial insights, and frequent reminders helped
complete this book.
The author also owes his students over the past 24 years a debt of
gratitude for their comments and questions which helped clarify so many
­topics/concepts. From his students, the author learned a lot; this book
would not have been possible without their input.
CHAPTER 1

Physicochemical
Treatment Processes

Hazardous waste sites may be contaminated with various toxic compounds,


and therefore to remove these hazardous compounds, environmental engi-
neers need to develop an arsenal of technologies and the maturity to adopt
the most appropriate technology/technologies for a particular scenario.
These technologies may be broadly classified as physical–chemical (also
known as physicochemical), thermal, stabilization, and encapsulation of
the waste (so the toxic components cannot migrate out of it), biological, and
land disposal. This chapter discusses four common physicochemical tech-
nologies that are widely employed in hazardous waste decontamination.

1.1  AIR STRIPPING

Air stripping is a physicochemical process used to remove volatile com-


pounds from a liquid phase (water) to a gas phase (air) by promoting
vigorous contact between the two phases. It is routinely used to treat
groundwater contaminated with volatile organic compounds (VOCs) at
low concentrations.
The most common method of contacting air with contaminated water
is in a stripping tower, as shown in Figure 1.1. The tower is filled with
a high surface-area-to-volume-ratio packing material and air flows in
opposite direction to that of water, known as countercurrent flow. Mass
exchange takes place between the aqueous and the vapor phase, and the air
stream that leaves the tower at the top carries the contaminant with it, typ-
ically at a much lower concentration that meets air quality standards. The
treated groundwater can then be discharged to a surface water-receiving
body or pumped to effect groundwater recharge. If the concentration of
the volatile organic contaminant is too high in the influent, its resulting
2
Figure 1.1.  Sketch of an air-stripping tower (packing material image courtesy:
www.ceramic-honeycombs.com/plastic/Plastic_Tripack_Tower_Packing.htm)
Physicochemical Treatment Processes  •   3

concentration in the exhaust may be too high for air quality standards, and
thus the off-gases may need to be treated.
The design of an air stripper entails expedited mass transfer of the vol-
atile compound from the aqueous phase to the gas phase. The “Two-film”
theory (Whitman 1923) (a full discussion of this topic is outside the pur-
view of this book) states that the mass transfer of the volatile compound
between water and gas is limited by diffusion through thin films at the
water–gas interface. A mass balance for an air stripper is discussed in the
next section.

1.1.1  MASS BALANCE FOR AIR STRIPPER

where
 L3 
Qw = water flow rate   
T 

 L3 
Qa = air flow rate   
T 

M 
Cin = influent water conc.   3 
L 
4  •   HAZARDOUS WASTE MANAGEMENT, VOLUME II

M 
Cout = effluent water conc.  3 
L 

M 
Gin = influent air conc.   3 
L 

M 
Gout = effluent air conc.   3 
L 

(set by environmental regulations)

z = height above bottom of air stripper [ L ]
C(z) = water conc. at z  M3 
 L 

M 
G ( z ) =   air conc. at   z    3 
L 

Mass balance between O and Z:

Mass in [ Qw  C ( z ) +   QaGin   ] =  Mass out [Qw  Cout +   Qa  G ( z ) ] (1.1)

Assume that the air coming into the packed-bed tower is clean and free
of the contaminant that is being removed from the water, that is, Gin = 0:
Qw
G ( z) =  (C ( z ) −  Cout ) . (1.2)
Qa

For the overall air stripper column,

Q 
Gout =  w  (Cin −  Cout ) . (1.3)
 Qa 

With equilibrium per Henry’s law,


Qw
Gout = H ′ Cin =  (Cin − Cout ) , (1.4)
Qa

where H′ = dimensionless Henry’s constant (see Volume 1 for details of


this equilibrium constant) defines the minimum air to water flow rate ratio

 Qa  C −  Cout 1
 Q  = in  ≈ . (1.5)
w min H ′C in H ′

Qa
Stripping factor, S ≡ H′ (1.6)
Qw
Physicochemical Treatment Processes  •   5

= number of minimum air to water ratios
needed  for high-efficiency stripping

In an ideal case, S = 1
Practically, S = 2 to 10, 3.5 is optimal
If S < 1, air stripper cannot achieve desired removal
The required air stripper height is a function of mass transfer kinetics:
Mass balance for differential element of length ΔZ at height Z inside
the tower:
dm′
Qc ( z + ∆z ) − Qc ( z ) =  a  ∆V (1.7)
dt
Contaminant Mass in
mass in water water
inflow outflow
dm′  M 
= mass flux per unit area across air-water interface   2 
dt L T 

 L2 
a = interface area per unit volume of tower   3 
L 

∆V = volume in differential element   L3  =   AT ∆z

AT = cross-section area of tower   L2 

L3 M L3 M M L2 M
Check units = × 3 − × 3 = 2 × 3 × L3 =
T L T L LT L T
From thin-film theory with water-side control:
G ( z)
dm′ −  C ( z )
=   − Dw H ′ . (1.8)
dt δw

=
Dw
Sw
( )
  C ( z ) −  Ceq ( z ) . (1.9)


( )
=   K L   C ( z ) −  Ceq ( z ) .
(1.10)
L
where K L = liquid-phase mass transfer coefficient  
T 
G
Ceq = water conc. in equilibrium with gas conc. =
H′
6  •   HAZARDOUS WASTE MANAGEMENT, VOLUME II

Back to mass balance:

( )
Qw C ( z + Dz ) −   Qw C ( z ) =   K L C ( z ) −  Ceq ( z ) a   AT Dz, (1.11)

Qw C ( z + Dz ) −  C ( z )
 ×  = C ( z ) −  Ceq ( z ) . (1.12)
AT K L a Dz

Qw dc
In the limit as  Dz → 0    ×  = C ( z ) −  Ceq ( z ) , (1.13)
AT K L a dz

Qw dc
 ×  = dz, (1.14)
AT K L a C ( z ) −  Ceq ( z )

Qw Cin dc L

AT K L a Cout c − ceq
=   ∫ dz = L = the required tower height.
0
(1.15)

Qw
Again,  from equation 1.2, G ( z ) =  (C ( z ) −  Cout ) .
Qa
Qw
 [C ( z ) − C (out)]
G( z ) Q
Therefore, C( eq )  ( z ) =   =  a   (1.16)
H′ H′

Qw Cin dC
AT K L a ∫Cout
and L =     .
 Qw   C −  Cout   (1.17)
C −    H′ 
 Qa  

  Qw  
 
   C +  (C −  C )  Qa  
in  
   in out
 H′  
Qw  1    
L =  
AT K L a  Q
   ln    . (1.18)
w   Cout 
 Qa   
 1 −  
H ′   
 
 
Qa
Recall that stripping factor S = Qw
H′
  Cin  
 1 +    ( S − 1) 
Qw  S    Cout   = HTU × NTU ,
∴ L =    ln  (1.19)
AT K L a S − 1 S 
 
 
Physicochemical Treatment Processes  •   7

Qw
where HTU = height of transfer unit = and
AT K L a

  Cin  
 1 +    ( S − 1) 
 S    Cout  .
NTU = number of transfer units =   ln
S − 1  S 
 
 

The following example provides a step-by-step approach for the


­design of an air-stripping column.

Example 1.1 Air Stripping

An air-stripping column is to be designed for removal of chloroform


(CHCl3) from groundwater. Pumping rate = 1.5 MGD = 65.7 L/s. The
influent CHCl3 concentration is 200 mg/L and the effluent should be
≤50 mg/L. The water temperature is 20°C.
Find

a. Packing tower height (m)


b. Packing tower diameter (m)
c. Air flow rate (Qa, m3/s)

Solution

Required depth of packing in the column

Z = HTU × NTU

  Cin   
   ( R − 1)  + 1 
R  Cout  
NTU =   ln   

R−1 R
 
 

Qa
R = Stripping factor = H ′ (dimensionless)
Qw
Practical range of R = 2 – 10

Choose R = 3
  200   
3   50  (3  −  1)   +  1 
NTU =     ln    
3  −  1  3 
 
 
8  •   HAZARDOUS WASTE MANAGEMENT, VOLUME II

3  ( 4  ×  2 )  +  1  3
=   ln
2  3  = 2   ln [ 3]

= 1.65
Qa
Now R = H ′
Qw
H
H ′ =
RT
for CHCl3 H = ?
Use the equation

H = exp (A − B/T )

where A and B are regression coefficients and T is the temperature in


­Kelvin. H has the units “atm-m3/mol”
for CHCl3 A = 9.84; B = 4.61 × 103 (from LaGrega et al. 2010)

 4.61  ×  103 
H = exp  9.84 −  = 2.76 × 10 atm-m /mol
−3 3
 293
H
H ′ = dimensionless Henry’s constant =
RT
atm { m 3
2.76  ×  10−3
= mole
atm { m 3
8.205  ×  10−5   ×  293 kelvin
mole { kelvin
= 1.147 × 10−1 (dimensionless)
Qa
R = stripping factor = H ′
Qw
RQw
Qa =
H′
L
(3)  65.7  
 s  = 1.719 × 103 L/s = 1.719 m3/s
=
1.147  ×  10−1

 lb 
L = liquid loading rate  2 
 ft { s 
 lb 
G = air loading rate  2 
 ft { s 
If diameter of air-stripping column = D (ft)
Physicochemical Treatment Processes  •   9

L = liquid loading rate

= 65.7 L /s × 1 kg/L × 1,000 g/kg × 1lb/453.6 gm × 1/πD2/4 ft2

 lb 
= 184.4/D2  2 
 ft { s 
G = air loading rate
m3 kg g lb
1.719    ×  1.205  3   ×  1,000    ×  
s m kg 453.6 g
=
π D2
4
5.814 lb
=
D 2 ft 2 { s
184.4
L  ρa  D 2  1.205  0.5
  0.5 = 5.814  1,000 
G  ρw 
2
D

= 1.1

As a general practice, the pressure drop is maintained in the range of


0.25 to 0.5 in H2O per foot of tower to prevent flooding. From Figure 1.2,
to maintain the recommended range of pressure drop,

G2 F
= 0.01
ρ Aρw g

Use Berl Saddle packing size 13 mm (0.5 in.).


From Table 5.28 of Perry’s Chemical Engineers’ Handbook, 7th
edition:

F = 240
a = 466

0.1ρ A ρ w g
G=
F

0.1( 7.52  ×  10−2 )  ×  62.4  ×  32.2


=
240
10  •   HAZARDOUS WASTE MANAGEMENT, VOLUME II

Figure 1.2.  Pressure–drop correlation. Reproduced from Eckert (1970) with


permission

lb 5.814
= 7.93 × 10−2 =
ft 2 -s D2

5.814
D=
7.93  ×  10−2

= 8.6 ft ≈ 9 ft

Provide a diameter of 9 ft = 2.75 m


To find KLa, use the following correlation,
0.5
 m 
KLa = α × DL × (0.305 L/µ)1−n 
 ρ DL 
where α = packing constant = 490 (from Table 5.28 of Perry’s Chemical
Engineers’ Handbook, 7th edition)
Physicochemical Treatment Processes  •   11

DL = diffusion coefficient, m2/s

= 1.44 × 10−9 m2/s for CHCl3

kg
L = liquid mass loading rate,
m 2 -s

65.7  L × 1 kg
=
s × L × π ( 2.75)
2

4
kg
= 11.06
m 2 -s

µ = viscosity of water = 1.002 × 10−3 Pa-s


ρ = density of water = 1,000 kg/m3
n = 0.2 (from Table 5.28 of Perry’s Chemical Engineers’ Handbook, 7th
edition)

1 − 0.28
0.305  ×  11.06 
KLa = 490 × 1.44 × 10−9 × 
 1.002  ×  10−3 
0.5
 1.002  ×  10−3 
× 
 1,000  ×  1.44  ×  10−9 

= 7.056 × 10−7 × 3.464 × 102 × 26.38

= 6.45 × 10−3 (s−1)


L
HTU =
M w KLa

L = liquid molar loading rate kmole


m 2 -s
MW = molar density of water = 55.6 kmole/m3

L = 65.7 L/s × 1 kg/L × kmole/18 kg × 1


π ( 2.75)
2
m2
4
12  •   HAZARDOUS WASTE MANAGEMENT, VOLUME II

= 6.145 × 10−1 kmole


m 2 -s

6.145  ×  10−1
HTU =  m
55. 6  ×  6.45  ×  10−3

= 1.71 m

Tower height = HTU × NTU

= 1.71 m × 1.65

= 2.82 m ≈ 3 m

Thus, a column with a diameter of 9.0 ft and a height of 10 ft, packed


with 0.5 in. Berl saddle packing, which introduces air from the bottom at
a rate of 7.93 × 10−2 lb , can treat 1.5 MGD of a groundwater source
ft 2 -s
and reduce the CHCl3 concentration from 200 mg/L to ≤ 50 mg/L.

1.2 ADSORPTION

Adsorption of a substance involves its accumulation at the interface


between two phases. For hazardous waste treatment, the two phases
­
are liquid (water) and solid. The molecule that accumulates at the interface
is called the adsorbate, and the solid phase on which this accumulation
­(adsorption) occurs is called the adsorbent. In hazardous waste treatment,
adsorbates typically include toxic organic compounds that are produced
by chemical industries. Details of the types of organic hazardous com-
pounds are available in Volume 1 of this series. The most common adsor-
bent is activated carbon.
Activated carbon is manufactured from natural, carbonaceous mate-
rials such as coal, peat, and coconuts by several inexpensive (e.g., high
temperatures ∼800°C and steam) and proprietary processes. The most
basic scheme involves the pyrolytic carbonization and subsequent activa-
tion. During the carbonization step, volatile components are released from
Physicochemical Treatment Processes  •   13

the raw carbonaceous material and graphite is formed. In this step, the
carbon is realigned with the result that a porous network is formed. In the
activation step, carbon is removed selectively from an opening of closed
porosity and the average size of the micropores is increased. Typically,
the process of activation results in a product that has very high surface
area (>1,000 m2/g) and pore volume (>0.5 mL/g), and can result in high
adsorption capacity (see examples 1.2 and 1.3).
Activated carbon is available in essentially two particle size ranges:
Powdered Activated Carbon (PAC, mean particle size 20 to 50 μm) and
Granular Activated carbon (GAC, mean particle size 0.5 to 3 mm). For
more information on the principal uses, advantages, and disadvantages of
using PAC versus GAC, please see “Water Treatment: Theory and Prac-
tice” by Sengupta, which is a part of this series.

1.2.1  ADSORPTION ISOTHERM

The amount of adsorbate adsorbed corresponding to a particular equilib-


rium aqueous-phase concentration is termed an “adsorption isotherm” and
is widely used in hazardous waste treatment. Adsorption isotherms are
performed by contacting a known mass of adsorbate in a fixed volume
of liquid to various dosages of adsorbent. At the end of the equilibration
period, which typically lasts a week, the aqueous-phase concentration of
the adsorbate is measured and the adsorption equilibrium capacity is cal-
culated for each bottle using the mass balance expression:

V
qe =    (C0 −  Ce ) , (1.20)
M

where
qe = equilibrium adsorbent-phase concentration of adsorbate, mg adsorbate/g
adsorbent
C0 = initial aqueous-phase concentration of adsorbate, mg/L
Ce = equilibrium aqueous-phase concentration of adsorbate, mg/L
V = volume of aqueous phase added to bottle, L
M = mass of adsorbent, g
Two common equations used to describe the equilibrium relationship
are the Langmuir and the Freundlich models.
14  •   HAZARDOUS WASTE MANAGEMENT, VOLUME II

1.2.1.1  Langmuir Model

In the Langmuir model, the adsorbent surface is made up of fixed individ-


ual sites where molecules of adsorbate may be chemically bound. More-
over, the model allows accumulation only up to a monolayer.
The following equation shows the reversible reaction between a ­vacant
surface site, adsorbate molecule, and the adsorbate-surface site-bound
species:

Sv + A  ↔ SA,
(1.21)

where
Sv = vacant surface sites, mol/m2
A = adsorbate species A in solution, mol
SA = adsorbate species bound to surface sites, mol/m2
For the reaction shown in equation (1.21), the equilibrium expression is
[ SA]
K ad =   ,
[ v ] [ A]
S
(1.22)

where CA = equilibrium concentration of adsorbate A in solution, mol/L


Equation (1.22) is combined with the following equation:

ST =   Sv +   S A , (1.23)

where ST = total number of sites available for monolayer coverage, mol/m2


Simple algebraic manipulation of equations (1.22) and (1.23) results in

ST K ad  C A   ST
SA =   =  . (1.24)
 1  1 +   K ad  C A
1 + 
 K ad  C A 

Equation (1.24) is not very useful in practical application. However,


if each side is multiplied by the surface area of the adsorbent and the
­molecular weight of the adsorbate, it results in

K ad  C A   ST   Aad   MW Q K  C Q   b  C
qA = ( S A ) ( Aad ) ( MW ) =   =   m ad A =   m A A ,
1 +   K ad  C A 1 +   K ad  C A 1 +   bA   C A
(1.25)
Physicochemical Treatment Processes  •   15

where
qA = equilibrium adsorbent-phase concentration of adsorbate A, mg
­adsorbate/g adsorbent
Aad = surface area per gram of adsorbent, m2/g
MW = molecular weight of adsorbate, g/mol
CA = equilibrium concentration of adsorbate A in solution, mg/L
Qm = maximum adsorbent-phase concentration of adsorbate when s­ urface
sites are saturated with adsorbate, ST × Aad × MW, mg adsorbate/g
adsorbent
bA = Langmuir adsorption constant of adsorbate A, Kad, L/mg
Equation (1.25) can be more conveniently expressed as

CA 1 C
=    +   A . (1.26)
qA bA   Qm Qm

Equation (1.26) is the equation of a straight line. A plot of CA /qA


­versus CA results in a straight line with slope of 1/Qm and intercept of
1/(bA × Qm). Example 1.2 illustrates how the Langmuir model can be used
in a practical application.

1.2.1.2  Freundlich Model

The Freundlich model is used to quantify the adsorption of a wide range


of hazardous waste compounds on a heterogeneous adsorbent such as
­activated carbon. It has the form

qA =   K AC 1/ n
A , (1.27)

where KA = Freundlich adsorption capacity parameter, (mg/g)(L/mg)1/n


1/n = Freundlich adsorption intensity parameter, unitless
CA = equilibrium concentration of adsorbate A in solution, mg/L
Equation (1.27) can be easily linearized as

1
log ( qA ) = log ( K A ) +   log (C A ) (1.28)
n

and a log–log plot of qA versus CA will result in a straight line.


16  •   HAZARDOUS WASTE MANAGEMENT, VOLUME II

The Freundlich model is consistent with adsorption theory of a het-


erogeneous adsorbent (e.g., Halsey and Taylor 1947; Weber and DiGiano
1996). The parameter KA is related primarily to the adsorption capacity
of the adsorbate, and (1/n) is a function of the adsorbent heterogeneity. A
(1/n) value of 1 indicates simple partitioning for an adsorbent with a uni-
form pore size and surface chemistry, while a value <1 (as with activated
carbon) typifies a broad distribution of adsorption site energies. Examples
1.2 and 1.3 illustrate how the Freundlich model can be used in a practical
application.

Example 1.2

A pharmaceutical manufacturer plans to install a new industrial produc-


tion process. The stream for the process is expected to have a concentra-
tion of 7.5 mg/L of a toxic compound at a flow rate of 9.45 × 10−4 m3/s.
The company plans to use activated carbon adsorption to remove the toxic
compound. The following adsorption data were obtained from the labora-
tory. The liquid volume in each flask was 200 mL.
Determine the Freundlich parameters KA and 1/n and the Langmuir
parameters Qm and bA for this system at a temperature of 23°C. Estimate
the amount of activated carbon required to remove the toxic contaminant
each year if the carbon is removed by an adsorption column that reaches
equilibrium with the carbon at the 7.5 mg/L concentration.

Mass of Initial Concentration Final


Adsorbent (mg) (mg/L) Concentration (mg/L)
98.6 25 0.04
58.1 25 0.11
26.3 25 0.49
15.7 25 1.2
8.8 25 3.2
2.9 25 10.2
0.8 25 19.7
Solution

Mass Concentration Concentration Adsorbent CA/qA Log (CA) Log (qA) Langmuir Freundlich
Adsorbent
mg mg/l CA (mg/L) qA (mg/g) (g/L) (mg/g) (mg/g)
98.6 25 0.04 50.63 0.00079 −1.39794 1.704398 17.16 49.87
58.1 25 0.11 85.68 0.001284 −0.95861 1.932879 46.23 85.48
26.3 25 0.49 186.39 0.002629 −0.3098 2.270418 185.21 189.44
15.7 25 1.2 303.18 0.003958 0.079181 2.481707 381.85 305.27
8.8 25 3.2 495.45 0.006459 0.50515 2.695004 704.45 514.75
2.9 25 10.2 1,020.69 0.009993 1.0086 3.008894 1,080.23 954.51
0.8 25 19.7 1,325.00 0.014868 1.294466 3.122216 1,224.19 1,355.38

17
18  •   HAZARDOUS WASTE MANAGEMENT, VOLUME II

Freudlich Fit
3.5
y = 0.5327x + 2.4425
3
LOG qA (mg/g)
R2 = 0.9992
2.5
2 Freudlich Fit
1.5
1
Linear
0.5 (Freudlich Fit)
0
–2 –1 0 1 2
LOG CA (mg/L)

Langmuir Fit
0.02
y = 0.0007x + 0.0023
0.015
CA/qA (g/L)

R2 = 0.9407
0.01 Langmuir Fit

0.005 Linear
(Langmuir Fit)
0
0 20 40
Equilibrium Adsorbate
Concentration (CA, mg/L)

The Langmuir plot yields,


CA
= 0.0023 + ( 0.0007 × C A )
qA
Comparing it to equation (1.26), it is seen that
1 1
0.0023 =     and 0.0007 =   .
(Qm bA ) Qm
Thus, Qm = (0.0007)−1 = 1428.6 mg of the toxic hazardous compound/g
of activated carbon and bA = 0.304 L/mg of the toxic contaminant.
The Freundlich plot yields

log ( qA ) = 2.4425 +  0.5327 log (C A ) .

Comparing it to equation (1.28), it is seen that

KA = 10^(2.4425) = 277 (mg toxic compound/gm activated carbon)


(L/mg toxic compound)0.5327

∴ 1/n = 0.5327
Physicochemical Treatment Processes  •   19

For the Langmuir fit,


If the activated carbon reaches equilibrium with 7.5 mg/L of the toxic
compound,

qA = 1428.6 × 0.304 × 7.5/(1 + (0.304 × 7.5))


= 993 mg toxic ­compound/gm activated carbon

Mass of toxic compound to be removed in one year = 7.5 mg/L ×


9.45E-04 m3/s × 1,000 L/m3 × 86,400 s/d × 365 d/yr = 2.235 × 108 mg
Thus, mass of activated carbon required = 2.235 × 108 mg/993 mg/g
activated carbon = 2.25 × 105 gm = 225 kg.
For the Freundlich fit,

qA = 277 × (7.5) 0.5327 = 810.3 mg of the toxic compound/gm activated


carbon

Mass of the toxic compound to be removed in one year = 7.5 mg/L ×


9.45E-04 m3/s × 1,000 L/m3 × 86,400 s/d × 365 d/yr = 2.235 × 108 mg
Thus, mass of activated carbon required = 2.235 × 108 mg/810.3 mg/g
activated carbon = 2.76 × 105 gm = 276 kg.

Example 1.3

Water generated by a chemical facility is found to be contaminated


with a very toxic compound 1,2-Dibromo-3-Chloro-propane, DBCP or
­Nemagon, which is a pesticide. Calculate the operating costs of reducing
the DBCP concentration from 10 mg/L to 1 mg/L with PAC treatment.
PAC costs 40 cents/lb. Report your answer in dollar or cents/m3 of waste-
water. Assume operating costs are solely due to PAC consumption and the
­Freundlich isotherm for DBCP is

qA = 53 CA0.47

where qA = mg DBCP/gm PAC

CA = mg DBCP/L water

Solution

For one m3 of contaminated water,


DBCP adsorbed = (10 − 1) mg/L × 1,000 L/m3 = 9,000 mg
Equilibrium loading = qA = 53 × CA0.47 = 53 mg DBCP/gm PAC
20  •   HAZARDOUS WASTE MANAGEMENT, VOLUME II

PAC required = 9,000 mg DBCP/53 mg DBCP/gm PAC = 169.81 gm


PAC = 0.374
Therefore, PAC cost = 0.374 lb × 40 ¢/lb = 15 ¢/m3 of wastewater
treated.

1.2.2  GAC CONTACTOR DESIGN USING THE BED DEPTH—


SERVICE TIME (BDST) ANALYSIS

Pilot studies are typically conducted for designing full-scale GAC con-
tactors in hazardous waste treatment applications. Pilot studies are pre-
ferred over other approaches (theoretical and/or model based) because it
is very common to encounter hazardous waste treatment scenarios where
the waste is a mixture of many contaminants. Even if individual contam-
inants are identified and their concentration determined, which in itself
could be prohibitively time-consuming and expansive, the interactions
between components (of the mixture) can be extremely complicated and
it may well-nigh be impossible to predict with theoretical constructs or
modeling studies. In a typical pilot study, the hazardous waste solution is
pumped through a series of columns, each typically 5 cm in diameter and
2 to 3 m deep.
The contaminant concentration is monitored as a function of time,
and data for fractional breakthrough (the ratio of CCout ) is plotted versus
in
time, as shown in Figure 1.3.

Figure 1.3.  Pilot-scale studies: effluent profile for three columns


in series
Physicochemical Treatment Processes  •   21

Note that in Figure 1.3, three pilot-scale columns were placed in


s­ eries and the effluent from each was analyzed and plotted. Next, a hor-
izontal line is drawn through each of the three curves at CCoutin = 0.1 and
Cout
Cin = 0.9
. CCoutin = 0.1 represents the breakthrough point and CCoutin = 0.9
is considered the exhaustion. The horizontal line difference between
the exhaustion line and the breakthrough line is defined as the height of
the adsorption zone (D). The CCoutin = 0.9 and CCoutin =  0.1, known as the ser-
vice time, are each p­ lotted as a function of bed depth and fit to a straight
line by linear regression:

t = ax + b, (1.29)

where
 days 
a = slope  
 meter 

x = depth in column ( meter )

F1   N
a= (1.30)
Cin  V

F1 = 1,000 ( for SI units )

mass of containment removed  Kg 
N = adsorptive capacity for GAC =  3 
volume of GAC m

 mg 
Cin = influent contaminant concentration  
 L 

 m3 
 
V = superficial velocity through the column   h2 
 m 
 

 F2  C 
b = intercept =     ×  ln  in − 1 (1.31)
 K × Cin   Cout 

F2 = 1,000 ( for SI units )
22  •   HAZARDOUS WASTE MANAGEMENT, VOLUME II

K = adsorption rate constant required to move an adsorption 
zone through the critical depth

 m 3  of hazardous waste 
=
 kg GAC × hour 

 mg 
Cout = contaminant concentration at breakthrough  
 L 


GAC  utilization rate = × ( cross-sectional area,  m 2 )
 day 

 meter 
 kg 
×  bulk density, 3 
 m 
If GAC columns are placed in series, the total number of columns is
AZ
n= + 1 (1.32)
d
where
n = number of columns in series
AZ = height of the adsorption zone (m)
d = height of a single column (m)
Example 1.4 illustrates the approach of the BDST method.

Example 1.4

A groundwater plume is contaminated with a mixture of ­hazardous


compounds. The groundwater is treated by GAC adsorption at a rate of
gal
100,000 day . The treatment process reduces the total concentration of
mg mg
the hazardous waste from 100 L to 10 L ; that is, 90 percent reduc-
tion. P
­ ilot-scale studies are conducted in columns 2 m long and having
gal
a ­diameter of 5 cm. The setup is operated at a flow rate of 100 day . The
results are shown:

Column 1 Column 2 Column 3


t10 (days) 9 26 41
t90 (days) 25 43 58
Cumulative Bed Depth (m) 2 4 6

Determine: (i) the height of the adsorption zone (m)


(ii) the number and size of columns required
kg
(iii) the GAC usage rate day ( )
Physicochemical Treatment Processes  •   23

kg
Assume the bulk density of GAC = 480
m3  

From the graph above, it is found that the height of the adsorption
zone is 2.0 m.
2
Number of columns needed = + 1 = 2
2
2
π  5 cm  ×  m 
Area of one lab column =   ×   −3
 = 1.96 × 10  m
2
4  100 cm 

The loading rate in the pilot-scale setup


gal liters m3 1
= 100   ×  3.7854 × ×
day gal 1,000 L 1.96 × 10−3  m 2
m3
= 193.13 2
m { day
Using the same loading rate for the full-scale setup, the area of GAC
column required
gal liters m3 1 m 2 { day
= 100,000 × 3.7854 × ×
day gal 1,000 liters 193.13 m 3
= 1.96 m 2

π
× ( dia ) = 1.96 m 2
2
4

Dia required = 1.6 m


24  •   HAZARDOUS WASTE MANAGEMENT, VOLUME II

From the plot above,


days
Slope = a = 8
m
Intercept = b = −6.67 days
Equation of the Bohart–Adams (Bohart and Adams 1920)  line
= 8 x − 6.67
1 m
Velocity of the adsorption zone = = 0.125
a day
Carbon utilization rate
1 m Kg Kg
= area × × unit weight = 1.96 m 2 × 0.125  × 480  3 = 117.6 
a day m day

F1 N
a = 
Cin  V

a × Cin × V days mg m3 1,000 L × Kg


N = = 8  × 100 × 193.13 2 × 3
F1 m L m day m × 106  mg
8 × 100 × 193.13 kg
=
1,000 m 3
kg
= 154.5 3
m

That is, 1 m3 of GAC can remove 154.5 Kg of the contaminant.

1.3  ION EXCHANGE

Ion exchange is a reversible chemical reaction where, in the most com-


monly used form, a contaminant (undesirable) ion from solution is
exchanged with an equivalent amount of another (benign) ion of the same
charge, which is attached to an immobilized solid phase. The immobilized
solid-phase particles are either naturally occurring inorganic zeolites or
synthetic, organic ion exchange resins. The primary morphology of ion
exchangers is spheres of diameter 75 to 1,000 µm, but membranes, com-
posite sheets, and nanoparticles are increasingly being used for specialty
applications. The vast potential of this technology is derived from its abil-
ity to generate specific ion exchangers that have high affinity for target
ions. Thus, selectivity of the ion exchanger plays a critical role, and it
allows engineers to design ion exchange processes where the target ion
is almost completely removed from the influent, and the effluent con-
tains “nondetect” amounts of it. Moreover, ion exchange technology is
Physicochemical Treatment Processes  •   25

almost always employed as a cyclical process. Therefore, the ability of a


regenerating solution to desorb the target ion from the surface of the ion
­exchanger and allow reuse of the ion exchanger for another exhaustion
cycle is a critical factor in its success.

1.3.1  ION EXCHANGE THEORY

In an ion exchange reaction between the aqueous phase and the polymeric
ion exchanger, the latter can be viewed as a strongly ionized electrolyte
with a relatively low dielectric constant. Consequently, attempts have been
made to mimic specific aqueous-phase interactions in the solid phase.
The following interactions are often operative in selective ion exchange
processes:

1. Electrostatic (or coulombic type)


2. Hydrophobic (or Van der Waals type)
3. Lewis acid-base (metal ligand type)
4. Ion dipole
5. Dipole dipole
6. Ion sieving
7. Ion exclusion
8. Steric effect

Many of these interactions can be characterized and enhanced by


modifying or tailoring the following composition variables of the ion
exchangers:

• Polymer matrix
• Covalently attached functional groups
• Cross-linking
• Porosity

Also, in order for the ion exchange processes to be viable, the ion
exchangers should be amenable to regeneration or desorption so that
they may be used for hundreds of cycles. In fact, the overall economy of
an ion exchange process is often dictated by the operating costs of the
regenerant chemicals as opposed to the fixed cost of the ion e­ xchangers.
Thus, ideally, an ion exchanger process should be reversible so that the
target solutes can be desorbed efficiently, leading to energy-efficient
separation. In reality, however, efficiency of desorption (or regenera-
tion) tends to diminish for highly selective sorbents because of the
26  •   HAZARDOUS WASTE MANAGEMENT, VOLUME II

strong bonding (or high free energy change) between the target ion and
the sorbent.
In order to strike a balance between selectivity and regenerability,
the intensity of solute–sorbent interaction has to lie within an envelope
where ion exchange is selective, but at the same time reversible. Figure 1.4
helps quantify such a working regime for various types of interactions.
The importance of reversibility is also well recognized for homogeneous
separation processes using specialty complexing agents (King 1987). It
is worthwhile to note that more than one interaction may be operative for
the sorption (ion exchange) of a single solute. In such cases, individual
free energy changes are additive and may enhance the overall sorption
affinity. For example, perchlorate (a toxic anion) and chloride (a benign
anion) are identical electrostatically; that is, both are monovalent anions.
But perchlorate exhibits much higher affinity toward an anion exchanger
with polystyrene matrix and tributyl functional group due to accompany-
ing hydrophobic bonding between perchlorate and the resin matrix. Thus,
it is possible to synthesize an anion exchanger which is more selective
toward the toxic perchlorate anion compared to the benign chloride anion.

Figure 1.4.  A quantitative measure of various interactions in ion exchange type


sorption processes
Source: SenGupta (1995).
Physicochemical Treatment Processes  •   27

1.3.2  ION EXCHANGE EQUILIBRIUM

The preference of the ion exchanger for the target ion is often expressed as
a separation factor, αij or a selectivity coefficient Kij for binary exchange
(Clifford 1990). Consider, for example, the cation exchange reaction of
Na+ for H+ in a cation exchanger:

R − H + + Na + ↔ R − Na + + H + (1.33)

{R − Na + } {H + }
K = (1.34)
{R − H + } {Na + }

In equations (1.33) and (1.34), overbars denote the ion exchanger phase.
For a cation-exchange reaction as shown above, the exchangeable ions (Na+
and H+) are termed counterions and the negatively charged ions in solution are
termed coions. K is the thermodynamic equilibrium constant, and {} denotes
ionic activity. Due to convenience of measurement, concentrations are used
in practice in place of activities. In this case, equation (1.35) based on con-
centration, the selectivity coefficient KNa/H ­describes the exchange. Note that
KNa/H includes activity coefficient terms that are functions of ionic strength
and, thus, is not a true constant, varying with different ionic strengths.

[ R − Na + ][ H + ] qNa CH
K Na / H = =
[ R − H + ][ Na + ] qHCNa (1.35)

where [ ] = concentration, mole/L


qNa = resin-phase concentration of sodium, equivalent/L
CNa = aqueous-phase concentration of sodium, equivalent/L
The binary separation factor αNa/H is a useful parameter that describes
the ion exchange equilibria since it is dimensionless:
distribution of ion i between phases y / xi
α ij = = i (1.36)
distribution of ion j between phases yj / xj
Thus, for the example above:

yNa / xNa y x ( q / q) (CH / CT ) qNa CH


α Na / H = = Na H = Na = (1.37)
yH / xH xNa yH (CNa / CT ) ( qH / q) qHCNa
where
yNa = equivalent fraction of Na in resin, qNa/q
yH = equivalent fraction of H in resin, qH/q
qNa = concentration of Na on resin, equivalent/L
28  •   HAZARDOUS WASTE MANAGEMENT, VOLUME II

qH = concentration of H on resin, equivalent/L


q = total exchange capacity of resin, equivalent/L
xNa = equivalent fraction of Na in water, CNa/CT
xH = equivalent fraction of H in water, CH/CT
CT = total ionic concentration of water, equivalent/L
Equations (1.35) and (1.37) show that for homovalent exchange, the
separation factor αij and the selectivity coefficient Kij are equal.
qNa CH
K Na / H = = α Na / H
qHCNa
However, for heterovalent ion exchange, the separation factor is not
equivalent to the selectivity coefficient. If the case of calcium-sodium
­exchange is considered,

2( R − Na + ) + Ca 2 + ↔ ( R − )2 Ca 2 + + 2 Na + (1.38)

2
qCa CNa
K Ca / Na =
qNaCCa (1.39)
2

yCa / xCa
while α Ca / Na = (1.40)
yNa / xNa

Equations (1.39) and (1.40) can be combined to obtain,

q / CT (1.41)
α Ca / Na = K Ca / Na
xNa / yNa

Equation (1.41) clearly shows that for heterovalent ion exchange the
separation factor can be manipulated through the total ionic concentra-
tion of the solution, CT. The capacity of the resin, q, is a constant. The
higher the CT, the lower the αCa/Na; that is, selectivity tends to reverse in
favor of the monovalent ion as the ionic strength increases. There is a
general selectivity sequence for a typical cation or anion exchanger as
is shown in Table 1.1. From Table 1.1, it can be understood that a sodi-
um-loaded cation exchanger can be used to remove a divalent cation such
as Ca2+ from a dilute concentration of water—as is the case with hard
water—since selectivity for Ca2+ > Na+. When the resin is exhausted, a
very high-concentration Na+ solution (as NaCl) is passed through the resin
bed. Based on equation (1.41), as the CT is increased, αCa/Na is decreased
and for a certain CT value, αCa/Na = 1.0. If CT is increased further, αCa/Na <
1.0; that is, selectivity reversal takes place and now the resin prefers Na+
to Ca2+, which means that Ca2+ is easily eluted from the ion exchanger and
it is back in Na+ form, where it can be used for another exhaustion cycle.
Physicochemical Treatment Processes  •   29

This is best explained by the ion exchange isotherm shown in


Figure 1.5. An ion exchange isotherm is a constant temperature equilib-
rium plot of ion exchanger-phase concentration versus aqueous-phase
concentration. A ­“favorable” isotherm (convex to the x-axis) means that
species i (which is plotted on each axis) is preferred to species j, the hid-
den or exchanging species. Thus, if the isotherm is a 45° line, it means that
the ion exchanger has no preference of i over j. And if the curve is concave
to the x-axis, it indicates that j is preferred over i. Example 1.5 shows how
to generate an ion-exchange isotherm.

Table 1.1  Relative selectivity of various counter-ions for a strong-acid


cation exchanger

Counterion Relative selectivity for AG 50W-X8 resin


H+ 1.0
Na+ 1.5
NH4+ 1.95
+
K 2.5
Cu+ 5.3
Ag+ 7.6
Mn2+ 2.35
2+
Mg 2.5
Fe2+ 2.55
Zn2+ 2.7
Co2+ 2.8
2+
Cu 2.9
Cd2+ 2.95
Ni2+ 3.0
Ca2+ 3.9
2+
Sr 4.95
Hg2+ 7.2
Pb2+ 7.5
Ba2+ 8.7
30  •   HAZARDOUS WASTE MANAGEMENT, VOLUME II

AG 50W-X8 is a strong-acid cation exchanger with sulfonic acid


functional group attached to a styrene-divinylbenzene copolymer matrix
with 8 percent divinylbenzene and is available from Biorad Inc., CA.

Figure 1.5.  Ion exchange isotherm for Ca2+ / Na+ exchange


showing selectivity reversal at high solution concentration

Example 1.5

An anion exchanger was preloaded with nitrate to saturation; that is, all the
anion exchange sites consisted of nitrate. This exchanger was equilibrated
with a 5 meq/L sulfate solution, and the results are provided in the table
below. Draw the sulfate/nitrate exchange isotherm for an anion concentra-
tion of 5 meq/L.
NO3− Initially NO3− on SO42− on
2− −
gm resin per Eq. [SO4 ] Eq. [NO3 ] xSO4 on Resin Resin at eq. Resin at eq. ySO4 αSo4/NO3
100 mL soln. meq/L meq/L meq meq meq

0.03 4.02 0.9 0.817073 0.09 0 0.098 1


0.10 2.15 2.6 0.452632 0.3 0.04 0.285 0.876923 8.616279
0.20 0.496 3.97 0.111061 0.6 0.203 0.4504 0.689317 17.7587
0.40 0.065 4.56 0.014054 1.2 0.744 0.4935 0.398788 46.5335
1.20 0.03 5.3 0.005629 3.6 3.07 0.497 0.139333 28.60043

Resin originally in nitrate form


Nitrate capacity of resin = 3.00 meq/g
Total solution concentration = 5 meq/L

31
32  •   HAZARDOUS WASTE MANAGEMENT, VOLUME II

1.3.3  MULTICOMPONENT EXCHANGE

Consider ions i . . . n in a solution in equilibrium with an ion exchanger.


The following assumptions are implied:

• Instantaneous mass transfer


• No shrinking or swelling of the resin
• Constant total exchange capacity
• Constant separation factors

At equilibrium,
Ci Cj C
xi = o
 ,  x j = o  ,  xk = ko (1.42)
,
C C C

qi qj q
yi =  ,  y j =  ,  yk = k . (1.43)
Q Q Q

Also, Ci + C j + Ck + ..... = C 0 (1.44)

xi + x j + xk + ..... = 1 (1.45)

qi + q j + qk + ..... = Q (1.46)

yi + y j + yk + ..... = 1 (1.47)

yi x j qi C j 1
Separation factor, α ij = = = (1.48)
y j xi q j Ci α ji

Now applying the conditions of equality in 1.48 into 1.46,


yi x j yi xk
yi + α ji + α ki + ..... = 1
xi xi

 xj x 
yi  1 + α ji + k α ki + ..... = 1
 xi xi 

1
yi =
xj
1+ ∑ α ji xi
j ≠i
Physicochemical Treatment Processes  •   33

xi (1.49)
yi =
xi + ∑ α ji x j
j ≠i
Similarly,

xj
yj = (1.50)
xj + ∑ α nj xn
n≠i

Thus, at equilibrium, the exchanger phase composition for a given


solution concentration can be determined for a multicomponent system
based on constant separation factor values.

Example 1.6 Fixed-Bed Process

A fixed-bed cation exchange column is considered to remove 10 mg/L of


Cu2+ from an industrial wastewater also containing 250 mg/L of Na+ and
100 mg/L of Ca2+ at pH 6.6. Consider the following two situations:

a. A cation exchange resin with sulfonic acid functional group that


has separation factor values of

αCu/Ca = 0.9, αCa/Na = 4.0

b. A chelating cation exchanger with iminodiacetate functional group


has separation factor values of

αCu/Ca = 75.0, αCa/Na = 8.0

Both resins have a capacity of 1.2 eq/L.

Questions

1. Compute yCu values for both resins at equilibrium


2. How many liters, or bed volumes (BVs), of wastewater can be
treated by one liter of each resin before copper breaks through from
the column? Show the effluent histories.
3. Do the anions present in the wastewater have any impact on the
results?

State assumptions if any.

Solutions

1. The equilibrium conditions are


34  •   HAZARDOUS WASTE MANAGEMENT, VOLUME II

Mass Equivalent
Concentration Concentration Equivalent
Name (mg/L) (meq/L) Fraction (xi)
Ca2+ 100 5 0.31
Cu2+ 10 0.32 0.02
Na+ 250 10.87 0.67

Using equations (1.49) and (1.50),

xCa
yCa =
xCa + α Cu ⋅ xCu + α Na ⋅ xNa
Ca Ca

xCu
yCu =
xCu + α Ca ⋅ xCa + α Na ⋅ α Ca ⋅ xNa
Cu Ca Cu

xNa
yNa =
xNa + α Ca / Na ⋅ xCa + α Cu / Ca ⋅ α Ca / Na ⋅ xCu

For strong acid cation (SAC) exchange resin with the following
properties

α Cu / Ca = 0.9, α Ca / Na = 4.0,  Q = 1.2  eq / L

yCa = 0.626, yCu = 0.036, yNa = 0.338

Check: yCa + yCu + yNa = 1.0


For iminodiacetate (IDA) chelating resin with the following properties

α Cu / Ca = 75.0, α Ca / Na = 8.0,  Q = 1.2  eq / L

yCa = 0.164, yCu = 0.80, yNa = 0.044

2. Sulfonic acid cation exchange resin has yCu = 0.023 at Q = 1.2


eq/L, or a capacity of 0.028 eq Cu2+/L. Iminodiacetate cation
exchange resin has yCu = 0.675 at Q = 1.2 eq/L, or 0.81 eq Cu2+/L.
With a feed of 0.16 meq/L, the resins can treat 175BVs (cationic)
and 5063BVs (iminodiacetate). Considering instantaneous break-
through with ideal plug-flow behavior, the breakthrough curves are
Physicochemical Treatment Processes  •   35

1.4  REVERSE OSMOSIS

Strictly speaking, reverse osmosis is not a treatment process because the


hazardous compounds are not transformed into harmless end products.
In this process, the waste stream containing dissolved solutes is forced
through a membrane under pressure. Ideally, all contaminants are trapped
by the pores of the membrane and only water molecules penetrate the
membrane, thus enabling separation of the contaminants from the water
phase. The pressure imposed on the feed solution has to be higher than the
osmotic pressure of the solutes. This forces the water molecules to move
through the semipermeable membrane, leaving the solutes behind, but not
completely. The primary goal of reverse osmosis is to reduce the concentra-
tion of these solutes in the product water. But the flip side is that the reject
stream has a much higher concentration of the solute. The reject stream,
however, has much lower volume compared to the treated water. This con-
centrated reject stream has to be treated before it can be discharged.
Principles of osmosis need to be understood before discussing Reverse
osmosis (RO). As shown in Figure 1.6(a), if a pure water solution and a
solution with a high concentration of solute (the hazardous compounds)
with free surface at the same elevation are separated by a semipermeable
membrane which allows only water molecules to pass, then the chemical
potential of the solute creates a disturbance to the equilibrium. To move
the system back to equilibrium, pure water moves to the solution contain-
ing the solute (at high concentration) to decrease the chemical potential of
the latter chamber. This process continues until an osmotic head (pressure)
is developed in the solution containing the solute, as shown in Figure 1.6
(b). Osmotic pressure of a salt solution across a semipermeable membrane
is computed by the Van’t Hoff equation:

π =  ϕ c  N  Cs  R  T , (1.51)
36  •   HAZARDOUS WASTE MANAGEMENT, VOLUME II

where π = osmotic pressure (atm)


ϕc = osmotic pressure coefficient
N = number of ions in each molecule
Cs = salt concentration (gmol/L)
R = universal gas constant = 0.082 (atm.L/(gmol.K)
T = absolute temperature (K)
If mechanical pressure much higher than osmotic head is applied on
the solute-solution side, water molecules move from the solute chamber to
the pure water side, as seen in Figure 1.6 (c).
Other important relationships include
Qp
Y = 100 ×   , (1.52)
Qf
where Y = Yield (%)
Qp = flow rate of permeate; that is, treated water
Qf = flow rate of feed; that is, influent water
Cp
S p = 100 ×   ,  (1.53)
Cf
where Sp = Solute passage (%)

Figure 1.6.  Flow across a semipermeable membrane caused by osmotic pressure


and (a) when pressure > osmotic pressure is applied to the solute (c)
Physicochemical Treatment Processes  •   37

Cp = solute concentration of permeate (mg/L or gmol/cm3)


Cf = solute concentration of feed (mg/L or gmol/cm3)
Also, Sr = solute rejection (%) = 100 − Sp
The flux of water and solute through the semipermeable membrane
are given as

J w =  W p ×  ( DP −   Dπ ) , (1.54)

( )
J s   =   K p ×   C f −  C p , (1.55)

where Jw = water flux through the membrane [gmol/(cm2.s)]


Wp = coefficient of water permeation [gmol/(cm2.s.atm)]
Δ P = pressure differential across membrane = Pin − Pout (atm)
Δ π = osmotic pressure differential across membrane = πin − πout (atm)
Js = solute flux through the membrane [gmol/(cm2.s)]
Kp = coefficient of salt permeation (cm/s)
Example 1.7 illustrates how all this information is used to design an
RO system.

Example 1.7

A wastewater stream at 25°C is to be treated by a reverse osmosis system.


The wastewater has a flow rate of 100 gal/min and the Total Dissolved ­Solids
(TDS, a measure of combined concentration of hazardous substances and non-­
hazardous benign compounds) is 10,000 mg/L. A 75 ­percent yield is required.
An RO vendor has recommended a membrane with the following
specifications:

Wp = 1.9 × 10-6 gmol/(cm2.s.atm)

Area of each bundle = 500 ft2


50 percent recovery rate
Osmotic pressure differential across the membrane = 600 psi
95 percent rejection of the salt

Determine:

1. The water flux through the membrane


2. Number of units required for 75 percent recovery
3. The final product water quantity and quality
4. The reject water quantity and quality
38  •   HAZARDOUS WASTE MANAGEMENT, VOLUME II

Solution

It should be noted that RO cannot selectively remove the hazardous com-


pounds. It will reject all compounds (benign and hazardous) to the same
degree.
A rule of thumb is that the osmotic pressure of a salt is 10 psi for
every 1,000 mg/L TDS. Since the TDS is 10,000 mg/L in this case, it is
assumed that the osmotic pressure differential is 100 psi.
The water flux through the membrane is given as Jw = Wp × (ΔP − Δπ)

Δ π = 100 psi (from rule of thumb)

gmol 1 atm
Jw = 1.9 × 10−6 (600 psi − 100.0 psi)
(cm 2 .s.atm) 14.7 psi

= 6.46 × 10−5 [gmol/(cm2.s)]

To determine the number of units, a schematic diagram of flow across


the membrane is drawn,

Since the membrane achieves 50 percent water recovery, the per-


meate flow rate (Qp) will be 50 percent of the feed flow rate (Qf); that
is, 50 percent of 100 gpm = 50 gpm. The reject flow rate (Qr) will be
the difference of flow rate of the feed water and the permeate; that is,
100 – 50 = 50 gpm. It is also necessary to do a mass balance on the
solute. It is given that the membrane rejects 95 percent of the salt;
therefore, the solute concentration of the permeate (Cp) = 5 percent
of the solute concentration of the feed (Cf); that is, 5 percent of 10,000
ppm = 500 ppm.

Moreover, Qf × Cf = [(Qp × Cp) + (Qr × Cr)]

On solving, Cr = 19,500 ppm


Physicochemical Treatment Processes  •   39

Thus, module 1 parameters are

To flow 50 gpm of permeate water through the membrane, membrane


area required
gal L 1 min 1,000 g gmol cm 2   ×  s
= 50  × 3.79  ×  ×   ×  × 
min gal 60 s L 18 g 5.24  ×  10−5  gmol

= 2.81 × 106 cm2

Each bundle of RO membrane has an area of 500 ft2. Therefore, the


number of bundles needed for module 1 is

2.81 ×  10  cm ×  


6 2 bundle
×  
( 3.28  ft ) ×   1 m 2 = 6.04
2

500 ft 2   m2 (100 cm )2  
Therefore 7 bundles are required for the first module.
Each pass has a 50 percent water recovery rate. In order to achieve
75 percent recovery, take the reject stream from pass 1 and use it as feed
stream to module # 2

Similar to pass 1, find the osmotic pressure of permeate in pass 2

100 psi
Osmotic pressure of feed = × 19,500 ppm = 195.0 psi
10,000 ppm
40  •   HAZARDOUS WASTE MANAGEMENT, VOLUME II

gmol 1 atm
Jw = 1.9 × 10−6 (600 psi – 195.0 psi)
2
(cm .s.atm) 14.7 psi

= 5.24 × 10−5 [gmol/(cm2.s)]

For pass 2, membrane area required

gal L 1 min 1,000 g


= 25  × 3.79  ×  ×   
min gal 60 s L
gmol cm 2   ×   s
×  × 
18 g 5.24  ×  10−5  gmol

= 1.67 × 106 cm2

Each bundle of RO membrane has an area of 500 ft2. Therefore, the


number of bundles needed for module 2 is

bundle ( 3.28 ft )
2
1 m 2
1.67 ×  106  cm 2 ×   ×   ×   = 3.6
500 ft 2   m2 (100 cm )2  
Therefore 4 bundles are required for the second module.
Thus, total number of bundles required for 75 percent recovery =
7 + 4 = 11
The final product water quality and quantity

Cp = 5 percent of Cf = 5 percent of 19,500 ppm = 975 ppm

Final water quantity = 50 gpm (from pass 1) + 25 gpm (from pass 2)


= 75 gpm
Final permeate salt concentration

( 50 gpm × 500 ppm ) + ( 25 gpm × 975 ppm ) 


= 
( 50 + 25) gpm
= 658.3 ppm

For pass 2, Qf = 50 gpm, Qp = 25 gpm, Qr = 25 gpm, Cf = 19,500


ppm, Cp = 975 ppm
Thus, Cr = 38,025 ppm
Therefore, the final reject quantity = 25 gpm
Final reject water quality = 38,025 ppm (almost four times as con-
centrated as the influent feed solution). This waste stream has to be treated
before disposal.
Physicochemical Treatment Processes  •   41

REFERENCES

Bohart, G. S., and E. Q. Adams. 1920. “Some Aspects of the Behavior of Char-
coal with Respect to Chlorine.” Journal of the American Chemical Society 42,
no. 3, pp. 523–44.
Clifford, D. 1990. Ion Exchange and Inorganic Adsorption in Water quality and
treatment: a handbook of community water supplies. American Water Works
Association. 1990. New York: McGraw-Hill.
Eckert, J. S. 1970. “Selecting the Proper Distillation Column Packing.” Chemical
Engineering Progress 66, no. 3, pp. 39–44.
Halsey, George, and Hugh S. Taylor. 1947. “The Adsorption of Hydrogen on Tungsten
Powders.” The journal of chemical physics 15, no. 9: pp. 624–30.
King, C. J. 1987. Chapter 15 in Handbook of Separation Process Technology,
R. W. Rousseau, ed, New York: John Wiley & Sons.
LaGrega, Michael D., Phillip L. Buckingham, and Jeffrey C. Evans. 2010. Hazardous
Waste Management. Long Grove, IL: Waveland Press.
SenGupta, A. K. 1995. Introduction of Ion Exchange Technology: Advances in
Pollution Control, Technomic Publishing Co., Inc., Lancaster, PA.
Weber Jr, W. J., and F. A. DiGiano. 1996. Process Dynamics in Environmental
Systems. New York, NY: John Wiley & Sons, Inc.
Whitman, W. G. 1923. “The Two-film Theory of Gas Absorption.” Chemical Met-
allurgical Engineering 29, pp. 146–48.
Index

A mass balance around separator,


Absorption, 56–57 87–89
Activated carbon, 12 solids retention time, concept
Adsorbate, 12 of, 86–87
Adsorbent, 12 Biotransformation, 63
Adsorption, 12–13
GAC contactor design using bed C
depth–service time (BDST) Cement, 57–58
analysis, 20–24 Competitive inhibition, 70–72
isotherm, 13–20 CSTR bioreactor design, at steady
Adsorption, 57 state, 79–82
AG 50W-X8, 30 biokinetic constants, 82–84
Air stripping, 1–3 mass balance
mass balance for, 3–12 around reactor, 84–85
American Society for Testing of around separator, 87–89
Materials (ASTM), 100 solids retention time, concept of,
86–87
B
Bed depth–service time (BDST) D
analysis, GAC contactor Data management plan, 104
design using, 20–24 Detoxification, 57
Biokinetics, 66–69 Double liner under waste,
constants, 82–84 92–93
inhibited reactions, 70–79 Drainage layer design, 95
Biological treatment, 63–64
batch systems, 64–65 E
biokinetics, 66–69 Eadie–Hofstee plot, 69
inhibited reactions, 70–79 Environmental Protection Agency
CSTR bioreactor design at (EPA), 59, 92
steady state, 79–82 EPA. See Environmental
biokinetic constants, 82–84 Protection Agency
mass balance around reactor, Extraction procedure toxicity
84–85 test, 60
108  •   INDEX

F L
“Favorable” isotherm, 29 Land disposal
Fixed-bed process, 33–35 landfills, 91–92
Flexible membrane liner (FML) cap requirements, 94–97
conductivity, 95–96 design and ­construction
design and construction of, of FML, important
­important considerations in, 97 ­considerations in, 97
Fluidized bed type, 49 double liner under waste,
FML. See Flexible membrane liner 92–93
Freundlich model, 15–20 leachate collection and removal
systems, 93–94
G leak detection requirements, 94
GAC contactor design, using bed Landfills, 91–92
depth–service time (BDST) cap requirements, 94–97
analysis, 20–24 design and construction of FML,
Glassmaking. See Vitrification important considerations
in, 97
H double liner under waste, 92–93
Halogens, wastes containing, 45–46 leachate collection and removal
systems, 93–94
I leak detection requirements, 94
Incineration, 50–51 Langmuir model, 14–15
basic reactions, 2.1 Leachability, 59
type of, 48–49 Leachant, 59
specific items on, 50 Leachate, 59
stoichiometric oxygen collection and removal systems,
­requirement, 51–54 93–94
wastes containing Leaching tests, 59
halogens and sulfur, 45–46 extraction procedure toxicity
nitrogen, 46–48 test, 60
Inhibited reactions liquid release test, 60
competitive inhibition, 70–72 paint filter test, 59
noncompetitive inhibition, 74–76 toxicity characteristics leaching
substrate inhibition, 76–79 procedure, 60–61
uncompetitive inhibition, 72–74 Leak detection requirements, 94
Inorganic agents Lime, 58
cement, 57–58 Lineweaver–Burk plot, 68
lime, 58 Liquid injection system, 49
pozzolans, 58 Liquid release test, 60
Ion exchange, 24–25
equilibrium, 27–32 M
multicomponent exchange, 32–35 Macroencapsulation, 56
theory, 25–26 Mass balance
Isotherm, adsorption around reactor, 84–85
Freundlich model, 15–20 around separator, 87–89
Langmuir model, 14–15 Mass balance, for air stripper, 3–12
INDEX  •   109

Mineralization, 63 S
Minimum Technical Requirements Sampling protocol, 100
(MTR), 92 Sintering, 57
Multicomponent exchange, 32–35 Site characterization, steps in,
99–104
N Soil component, 95
Nitrogen fixation, 47 Solids retention time, concept of,
Nitrogen, wastes containing, 46–48 86–87
Noncompetitive inhibition, 74–76 Stabilization and solidification
(S/S), 55–56
P agents, 57
Paint filter test, 59 inorganic agents, 57–58
Perfect combustion, 43 thermoplastic materials, 58
Personal protective equipment, vitrification, 58–59
101–103 leaching tests, 59
Physicochemical treatment extraction procedure toxicity
processes test, 60
adsorption, 12–13 liquid release test, 60
GAC contactor design using paint filter test, 59
the bed depth–service time toxicity characteristics leaching
(BDST) analysis, 20–24 procedure, 60–61
isotherm, 13–20 mechanisms, 56–57
air stripping, 1–3 Stoichiometric oxygen
mass balance for, 3–12 requirement, 51–54
ion exchange, 24–25 Substrate inhibition, 76–79
equilibrium, 27–32 Sulfur, wastes containing,
multicomponent exchange, 45–46
32–35
theory, 25–26 T
reverse osmosis, 35–40 TCLP. See Toxicity characteristics
Portland cement, 57 leaching procedure
Pozzolans, 58 Thermal VOC type, 49
Pre-entry offsite Thermoplastic materials, 58
characterization, 99 Toxicity characteristics leaching
Precipitation, 57 procedure (TCLP), 60–61

Q U
Quality assurance (QA) plan, 104 Uncompetitive inhibition, 72–74
Quality control (QC) plan, 104
V
R Vegetative cover, 94
Resource Conservation and Vitrification, 58–59
Recovery Act (RCRA) VOC. See Volatile organic
legislation, 92 compound
Reverse osmosis, 35–40 Volatile organic compound
Rotary kiln type, 49 (VOC), 48
OTHER TITLES IN OUR ENVIRONMENTAL
ENGINEERING COLLECTION
Francis J. Hopcroft, Wentworth Institute of Technology, Editor
• Environmental Engineering Dictionary of Technical Terms and Phrases: English to
Italian and Italian to English by Francis J. Hopcroft and Irena Stojkov
• Environmental Engineering Dictionary of Technical Terms and Phrases: English to
Vietnamese and Vietnamese to English by Francis J. Hopcroft and Minh N. Nguyen
• Environmental Engineering Dictionary of Technical Terms and Phrases: English to
Hungarian and Hungarian to English by Francis J. Hopcroft and Gergely Sirokman
• Management of Environmental Impacts by Alandra Kahl
• Protecting Clean Air: Preventing Pollution by Sarah J. Simon
• Environmental Engineering Dictionary of Technical Terms and Phrases: English to Greek
and Greek to English by Francis Hopcroft and Michos Georgios
• Cell and Molecular Biology for Environmental Engineers by Ryan Rogers
• Hazardous Waste Management, Volume I by Sukalyan Sengupta
• Technical Writing for Environmental Engineers by Joan Giblin and Emily Coolidge Toker
• Geology for Environmental Engineers by David Woodhouse and James Lambrechts
• Environmental Project Management by Cristina Cosma and Francis Hopcroft

Momentum Press offers over 30 collections including Aerospace, Biomedical, Civil,


Environmental, Nanomaterials, Geotechnical, and many others. We are a leading book
publisher in the field of engineering, mathematics, health, and applied sciences.

Momentum Press is actively seeking collection editors as well as authors. For more
information about becoming an MP author or collection editor, please visit
http://www.momentumpress.net/contact

Announcing Digital Content Crafted by Librarians


Concise e-books business students need for classroom and research

Momentum Press offers digital content as authoritative treatments of advanced engineering


topics by leaders in their field. Hosted on ebrary, MP provides practitioners, researchers, faculty,
and students in engineering, science, and industry with innovative electronic content in sensors
and controls engineering, advanced energy engineering, manufacturing, and materials science.

Momentum Press offers library-friendly terms:


• perpetual access for a one-time fee
• no subscriptions or access fees required
• unlimited concurrent usage permitted
• downloadable PDFs provided
• free MARC records included
• free trials

The Momentum Press digital library is very affordable, with no obligation to buy in future years.

For more information, please visit www.momentumpress.net/library or to set up a trial in the


US, please contact mpsales@globalepress.com.

You might also like