Transcript
Page 1: Permeable reactive barrier for groundwater remediation

www.elsevier.com/locate/jiec

Available online at www.sciencedirect.com

Chemistry 14 (2008) 145–156

Journal of Industrial and Engineering

Review

Permeable reactive barrier for groundwater remediation

R. Thiruvenkatachari a, S. Vigneswaran a,*, R. Naidu b

a Faculty of Engineering, University of Technology, Sydney, P.O. Box 123, Broadway, NSW 2007, Australiab CRC CARE, University of South Australia, Australia

Received 14 December 2006; accepted 26 October 2007

Abstract

This article aims to provide an overview of the upcoming technology of permeable reactive barriers for groundwater remediation. A

comprehensive list of references and web-links are also provided for further in-depth understanding. A brief discussion on the Australian

perspective on this emerging technology is also included.

# 2008 Published by Elsevier B.V. on behalf of The Korean Society of Industrial and Engineering Chemistry.

Keywords: Permeable reactive barriers; Groundwater; Remediation; Pollution

Contents

1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 145

2. Sources and types of groundwater contamination. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 146

3. Permeable reactive barrier. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 147

4. Configuration of PRBs . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 147

4.1. Conventional systems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 147

4.2. Advanced methods . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 147

5. Mechanism of interaction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 148

6. Treatment of inorganic and organic pollutants . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 149

7. Zero-valent iron. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 149

8. Activated carbon . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 150

9. Zeolites . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 150

10. Alkaline materials-complexing agents. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 150

11. Bioremediation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 151

11.1. Organic carbon for denitrification and sulphate reduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 152

12. Sequential reactive media . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 152

13. In situ chemical oxidation (ISCO) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 152

13.1. Use of oxidants . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 152

14. PRB studies in Australia and New Zealand. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 153

Acknowledgement . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 154

References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 154

* Corresponding author. Tel.: +61 2 9514 2641; fax: +61 2 9514 2633.

E-mail address: [email protected] (S. Vigneswaran).

1226-086X/$ – see front matter # 2008 Published by Elsevier B.V. on behalf of

doi:10.1016/j.jiec.2007.10.001

1. Introduction

Groundwater is a limited ecological resource representing a

small percentage of the total water distribution [1,2,3]. The

contribution from groundwater is vital; perhaps as many as two

billion people depend directly upon aquifers for drinking water,

and 40% of the world’s food is produced by irrigated agriculture

The Korean Society of Industrial and Engineering Chemistry.

Page 2: Permeable reactive barrier for groundwater remediation

R. Thiruvenkatachari et al. / Journal of Industrial and Engineering Chemistry 14 (2008) 145–156146

that relies largely on groundwater. Australia has 25,780 GL of

groundwater suitable for potable, stock and domestic use, and

irrigated agriculture that can be extracted sustainably each year.

It is extensively used for urban water supplies, agriculture,

irrigation, industry and mining. In Australia, some of the

regions like arid zones of South Australia, the Northern

Territory, and the Pilbara are entirely dependent on ground-

water. Most of the country’s premium wine districts rely on

groundwater. Due to the cap on surface water extractions in the

Murray–Darling Basin and the scarcity of surface water

resources in other areas, groundwater use across Australia has

increased significantly in the last 10 years [4]. South Australia,

New South Wales and Victoria use more than 60% of

groundwater for irrigation, while Western Australia uses

72% for urban and industrial purposes [5,6].

Studies by [4,7] identified increased demand for water in

Australia and called for proper management of groundwater.

The report also revealed that groundwater resource in Australia

has been highly committed in some places, or of poor quality in

others, and poorly investigated in others [8]. If not managed

properly, groundwater resources are highly vulnerable to

widespread contamination. There are many reports of serious

incidents of groundwater contamination due to accidental

spills, or unsatisfactory disposal of industrial chemicals,

agricultural practices, mining activities, etc.

Attempts at large-scale groundwater cleanup began in

earnest in the 1980s and the results of early remediation efforts

seldom produced the expected reduction in contamination

levels. Studies by the U.S. Environmental Protection Agency

(EPA) [9,10] found that the commonly used pump-and-treat

(P&T) technologies (pump the water and treat it at the surface)

rarely restored sites that had contaminated groundwater to

background conditions. This was confirmed in a much more

extensive 1994 National Research Council (NRC) study that

explicitly reviewed 77 sites across the United States where full-

scale pump-and-treat was being used [11,12]. One of the most

promising remediation technologies is the use of permeable

reactive barriers (PRBs) filled with reactive material(s) to

intercept and decontaminate plumes in the subsurface. In the

last decade, there has been an explosion of activities directed at

the development and implementation of PRBs. This study

presents a comprehensive review on PRBs technology.

2. Sources and types of groundwater contamination

Broadly, groundwater contaminants come from two cate-

gories of sources:

Table 1

Examples of point and diffuse pollutions

(a) P oint Sources and

Point source Non-point or diffuse pollution

(b) D istributed, or Non-Point Sources.

Municipal landfills, industrial waste

disposal sites, leaking gasoline storage

tanks, leaking septic tanks, and accidental

spills and leaks of petroleum products and

of dense industrial organics

Atmospheric deposition,

contaminated sediments,

and many land activities

that generate polluted runoff,

such as agriculture

(pesticides and fertilisers),

logging, and onsite sewage

disposal

Localised sources are known as point sources of contam-

ination. The contaminant interacts with the moving ground-

water and the soil and spreads out to form a plume moving in

the same direction as the groundwater. The resulting ground-

water contamination plume may extend several hundred metres

or even further away from the source of pollution.

Groundwater can also be contaminated by diffuse sources

over a wide area, for instance widespread use of fertilisers on

gardens and fields. Diffuse contamination may have greater

environmental impacts than contamination from point sources

because a much larger volume of water is affected. Pollutants

from point sources are generally related to urban development,

while diffuse sources are generally rural in nature. Some of the

examples of point and diffuse pollutions are given in Table 1.

Analysts estimate that there are between 300,000 and

400,000 sites in the USA contaminated with a wide variety of

toxic chemicals, representing clean up cost in the range of $500

billion to $1 trillion [11]. Many of these sites experience

groundwater contamination by complex mixtures of chlori-

nated solvents, fuels, metals, and/or radioactive materials.

Europe’s groundwater is polluted in several ways: nitrates,

pesticides, hydrocarbons, chlorinated hydrocarbons, sulphate,

phosphate and bacteria. Some of the most serious problems are

pollution by nitrates and pesticides. The key findings of

Australia: State of the Environment Report [13] highlighted that

Australia’s inland waters are under increasing pressure from

over-extraction, algal blooms, catchment modification, habitat

destruction and pollution. Also, the experiences from Europe

and North America suggest that groundwater pollution in

Australia will become a more serious issue in the future. There

are many well-documented cases of groundwater pollution in

Australia. The most significant diffuse contaminant of ground-

water throughout each state and territory in Australia is nitrates,

due to their adverse affects on people, animals and the

environment [14,15]. The main source of nitrate contamination

is through the application of fertilizers for cropping and pasture

[15]. Direct discharges of nitrogen compounds from on-site

sanitation and from sewer effluent also exacerbate the problem.

In many areas, the concentration is greater than the Australian

Drinking Water Guidelines [16] level of 50 mg/L nitrate (as

nitrate), resulting in groundwater that is unfit for drinking. In

some of the more contaminated areas, the concentration is in

excess of 100 mg/L [15].

Recent incidences of reported pesticide contamination of

groundwater in this country are listed in the [5] report. In most

affected areas, pesticides were detected in at least 20% of

samples, indicating significant contamination. However, sys-

tematic monitoring of pesticide contamination of groundwater in

Australia is limited indicating inadequate data on the quantities,

locations and types of pesticides used, as well as knowledge gaps

in the fate of pesticides in local environments [17].

Page 3: Permeable reactive barrier for groundwater remediation

R. Thiruvenkatachari et al. / Journal of Industrial and Engineering Chemistry 14 (2008) 145–156 147

3. Permeable reactive barrier

This technology termed as Permeable Reactive Barriers is

defined [9] as:

‘An emplacement of reactive media in the subsurface

designed to intercept a contaminated plume, provide a flow

path through the reactive media, and transform the

contaminant(s) into environmentally acceptable forms to

attain remediation concentration goals down-gradient of the

barrier’

The concept of PRBs is relatively simple. A permeable

reactive barrier material consisting of permanent, semi

permanent or replaceable reactive media is placed in the

subsurface across the flow of path of a plume of contaminated

groundwater, which must move through it as it flows, typically

under its natural gradient, thereby creating a passive treatment

system. As the contaminant moves through the material,

reaction occur that transform the contaminants into less

harmful (non-toxic) or immobile species. The PRB is not a

barrier to the groundwater, but it is a barrier to the

contaminants. PRBs are designed to be more permeable than

the surrounding aquifer materials so that contaminants are

treated as groundwater readily flows through without sig-

nificantly altering groundwater hydrogeology.

PRBs potentially have several advantages over conventional

pump-and-treat methods for groundwater remediation.

� P

RBs can degrade or immobilize contaminants in situ

without any need to bring them up to the surface. Hence no

need for expensive above ground facilities for storage,

treatment, transport, or disposal other than monitoring wells.

After the installation the above ground can be re-used for

other purposes. Also, as the contaminants are not brought to

the surface; there is no potential cross media contamination.

� T

hey also do not require continuous input of energy, because

a natural gradient of groundwater flow is used to carry

contaminants through the reactive zone. Only periodic

replacement or rejuvenation of the reaction medium might

be required after its reactive capacity is exhausted or it is

clogged by precipitants and/or microorganisms. However, the

drastically reduced operating costs offsets the higher

construction cost that are typical for PRBs, which results

in an overall reduction in the life cycle cost of this technology.

� D

egradation of most of the contaminants is achieved rather

than mere change of phase of contaminants. The barrier

provides effective contaminant remediation, much more than

simple migration control of the pollutants.

� T

echnical and regulatory problems related to ultimate

discharge requirements of effluent from pump-and-treat

systems are avoided with the PRB technology.

However, so far, limited data are available on the

performances of reactive barriers with different materials

and their comparative performances. Limited long-term field

testing data are available and field monitoring is in its infancy

[18].

4. Configuration of PRBs

4.1. Conventional systems

Two installation schemes are more frequently used in field

applications [19,20]; Continuous and Funnel-and-Gate PRB.

The continuous PRB configuration consists of a single

reactive zone installed across the contaminant plume, while the

funnel-and-grate system consist of a permeable gate (reactive

zone) placed between two impermeable walls that direct the

contaminated plume towards the reactive zone.

The choice between the two configurations depends on

both the hydrogeological characteristics of the site and the

reactive material cost [19]. When a high cost reactive material

is used, funnel-and gate configuration is preferred since the

reactive zone requires less material. However, construction

cost of continuous type barrier is much cheaper than funnel-

and-gate system. Hence a balance must be struck between the

cost of reactive material and the construction cost of the

barrier, in accordance with the target pollutant and level of

removal to be achieved. Multiple reactive medium in

succession or in series can be installed in the funnel-and

gate setup [20]. Alternatively, a relatively less expensive

method using multiple caisson gates can also be installed [20].

A caisson is a hallow, load-bearing (usually cylindrical)

enclosure generally used as an alternative method for

excavation. For the purpose of emplacing a reactive cell, a

prefabricated, steel caisson (normally 8-ft-dia or smaller) is

pushed or vibrated down into the subsurface. Once the caisson

has reached the intended depth, the soil within the caisson can

be augered out and replaced with the reactive medium. Upon

emplacement of the reactive cell/medium, the caisson can be

pulled straight out. The caisson can be installed from the

ground surface and completed without requiring personnel to

enter the excavation.

Usually, the conventional PRB installation techniques

require some degree of excavation, which limits the PRB to

fairly shallow depths of 20 m [21]. However, use of new

construction techniques, such as slurry injection and hydro-

fracturing are able to overcome this depth limitation.

4.2. Advanced methods

(i) Injection system: Injection system involves creating a

treatment zone within the contaminant boundary by

drilling series of bore holes or injection wells and injecting

the reactive material (chemical/particulate mixture) into

the treatment zone. Potential advantages of this approach

are that there is no need to construct a trench and possible

aquifer access at greater depths. Ususally, two or three

rows of overlapping, interlocking columns can offer

effective barrier [22]. Nevertheless, it has to be made sure

that the contaminant plume is efficiently taken care of and

no by-passing or fingering occurs, which may impair the

remediation effect.

(ii) H

ydraulic/pneumatic fracturing: Hydraulic/pneumatic

fracturing is intentional fracturing (cracking) of a subsur-

Page 4: Permeable reactive barrier for groundwater remediation

Table

Some

Grou

Orga

Me

Eth

Eth

Pro

Ar

Ot

Inorg

Tra

An

R. Thiruvenkatachari et al. / Journal of Industrial and Engineering Chemistry 14 (2008) 145–156148

face using pumped water and/or air under high pressure.

As the confining pressures are exerted in the borehole,

fractures will open and propagate out laterally from an

initiation point. A fracture fill slurry composed of the

reactive medium, can then be injected into the fracture to

form a reactive treatment zone. More than one fracture

may be required within the treatment zones. Fractures have

a preferred direction of propagation, they are therefore

asymmetric with respect to the borehole and they climb in

the preferred direction of propagation. Fractures can be

controlled to happen either horizontally or vertically

[24,25]. Some advantage to this technique includes the

ability to emplace a barrier to a depth greater than 80 ft.

Also, fracturing causes minimal disturbance, does not

generate contaminated soils, and is inexpensive. Fractured

zones may also be applied to direct groundwater flow

towards the gates in funnel-and-gate system [26]. Some

drawbacks of emplacement by hydraulic fracturing include

difficulty in controlling the fracture direction and the

limited soil conditions in which it can be used effectively.

(iii) P

assive groundwater capture and treatment by reactor

cells: This technique involves emplacement of reactor

cell(s) in the subsurface consisting of reactive medium and

capturing the contaminated plume into the reactor for

treatment [27]. This system does not involve any pumping

equipment and the plume is directed into the reactor by

siphoning or by natural gradient. Because of the passive-

mode of operation, the operating and maintenance

requirements are relatively minor.

The selection of the construction technique to be used

depends upon the site characteristics [25] such as depth of PRB,

geotechnical consideration, soil excavation: space for handling

and disposal of soil (contaminated), health and safety of

personnel. Common contaminants being treated using PRB

technologies are given in Table 2.

5. Mechanism of interaction

Proper understanding of the underlying process by which the

reactive material interact with the pollutants and the mechan-

ism of removal is very important. In general, the contaminant

2

of the common contaminants in groundwater for remediation

ps Pollutants

nic compounds

thane Tetrachloromethane, trichlo

anes Hexachloroethane, 1,1,1-tri

enes Tetrachloroethene, trichloro

panes 1,2,3-Trichloropropane, 1,2

omatics Benzene, toluene, ethylbenz

hers Hexaxhlorobutadiene, 1,2-d

anic compounds

ce metals/heavy metals Chromium, nickel, lead, ur

ion contaminants Sulphate, nitrate, phosphate

removal mechanisms can be classified broadly into three

categories [18]:

- D

ro

ch

et

-d

e

ib

an

,

egradation: Through chemical or biological reactions that

lead to decomposition or degradation of contaminants into

harmless compounds.

- P

recipitation: Immobilization of contaminants within the

reaction zone by formation of insoluble compounds. Here, the

chemical state of the contaminant is not altered.

- S

orption: Immobilization of contaminants within the reaction

zone by adsorption or complex formation. Here, the chemical

state of the contaminant is not altered.

More specifically, the principal processes are [29]:

� R

l

h

i

n

i

a

eductive degradation of organic pollutants.

� O

xidative degradation of organic pollutants.

� R

etardation and biodegradation of organic pollutants.

� S

orption of organic or inorganic pollutants.

� R

eduction and/or precipitation of heavy metal compounds.

In general, the types of reactive materials used for the

construction of permeable reactive barriers, are [30]:

� T

hose changing pH or redox potential,

� T

hose causing precipitation,

� M

aterials with high sorption capacity, and

� T

hose releasing nutrients/oxygen to enhance biological

degradation.

A study by USEPA [31], classified the reactive barrier

materials according to the target pollutant and the mechanism

of removal (Table 3).

Zero-valent iron (Fe0) is the most common reactive material

in the current field of application [19]. A comprehensive review

on iron in the use of PRBs has been carried out recently by [32].

Apart from iron based materials, other types of materials

[25,30,33] such as organic based (activated carbon, leaf, peat,

sewage sludge, sawdust, etc.) [34,35], alkaline- complexing

agents (hydrated lime, ferrous sulphate) [30,36], phosphate

minerals like hydroxyapatite and biogenic apatite (e.g. fish

bone) [37], zeolites [38–40], clay [41], metal oxides [38,42],

methane, dichloromethane

oroethane, 1,1,2-trichloroethane 1,1-dichloroethane

ene, cis-1,2-dichloroethene

chloropropane

e

romoethane, Freon 113, N-nitrosodimethylamine

um, iron, manganese, technetium, selenium, copper, cobalt, cadmium, zinc

rsenic

Page 5: Permeable reactive barrier for groundwater remediation

Table 3

Reactive materials classified based on the target pollutant and the mechanism of removal

Target and mechanism of removal Reactive materials

Inorganics-sorption or substitution

barriers

Activated carbon, activated alumina, bauxite, exchange resin, ferric oxides and oxyhydroxides, magnetite,

peat, humate, lignite, coal, phosphates, titanium dioxide, zeolite

Inorganics-precipitation barriers Biota, dithionite, ferrous hydroxides, ferrous carbonates, ferrous sulfide, hydrogen sulfide gas, lime, flyash,

limestone, miscellaneous (Mg(OH)2, MgCO3, CaCl2, CaSO4, BaCl2, zero-valent metals

Inorganics-degradation barriers Biota, zero-valent metals

Organics-degradation barriers Ferrous minerals, oxygen release, ultramicrobacteria, zero-valent metals

Organics-sorption barriers Zeolite, activated carbon, clays

R. Thiruvenkatachari et al. / Journal of Industrial and Engineering Chemistry 14 (2008) 145–156 149

microorganisms [43–46], polymers [47,48], innovative poly-

mer membrane (as liner) containing iron [49] are also being

evaluated.

Ref. [50] evaluated 124 PRB projects in the USA and

reported that in majority of projects (45%), zero-valent iron was

the material used as reactive barrier material. Another study

conducted by [51], on around 50 PRB sites across US, Canada

and Europe revealed that zero-valent iron is used in more

number of occasions. Refs. [52,53] reported that the primary

determinant of degradation rate in different in the case of iron is

the available reactive surface area. The parameter generally

used to discriminate between different irons is the specific

surface area, or the surface area per unit mass (m2/g) of iron. In

most cases, a linear relationship has been observed between

reduction rates and iron metal surface area concentration

[54,55]. Therefore commercial irons with higher surface area

are preferred [20]. Finer iron particles can also be mixed with

sand in order to increase the hydraulic conductivity. Iron with

several amendments has been tested to improve the perfor-

mance.

� D

uring the reaction, oxidation of Fe0 to Fe2+ results in an

increase in pH, which eventually leads to the formation of

precipitation of number of minerals. In order to overcome

this, pyrite or iron sulfide are added (oxidation of pyrite

produces acid) which lowers the pH and increases the organic

removal efficiency [56,57]. One negative effect of addint pH-

controlling amendments could be the increased possibility of

higher levels of dissolved iron in water downstream after the

reactive cell.

� I

nstead of using granular sized iron medium, use of colloidal-

size iron (1–3 mm) or of nanosize (1–100 nm = 0.001–

0.1 mm) would significantly increase the surface area and a

lower total iron mass may be required in the treatment zone

[20,58–60]. Nanosized [61,62] or as emulsion state [63] have

also been studied. Colloidal or nono-state iron allows the

formulation of slurries that can be injected in PRB by

boreholes or fractured media.

� B

imetallic systems where several metals are plated onto zero-

valent iron (ZVI) (e.g. Fe–Cu, Fe–Pd, Fe–NI) have shown to

produce greater performance compared to having ZVI alone

[64–66]. Some bimetals enhance the degradation by acting as

catalysts and as galvanic couples (increase electron activity).

However it has been cited that reactivity of bimetallic system

may be high initially but may decline gradually [20].

� A

cid pre-treatment of ZVI was also found to increase the

degradation efficiency; probably due to removal of oxide

coating layer on the iron surface or due to increase in surface

area by etching or pitting corrosion [52,67].

6. Treatment of inorganic and organic pollutants

Fundamental difference exists between organic and inor-

ganic contaminant remediation. Organic contaminants can be

broken down into innocuous elements and compounds, such as

carbon dioxide and water because they are molecules consisting

of carbon, hydrogen, halogens, oxygen, and sometimes sulphur,

phosphorous, and nitrogen. Conversely, most inorganic

contaminants are themselves elements. They cannot be

destroyed but can only change speciation [68]. Therefore

remediation strategies must focus on transforming inorganics

into forms that are non-toxic, not bioavailable, immobile, or

capable of being removed from the subsurface [68]. The

characteristics of these elements have in common is that they

can undergo redox reactions and can form solid precipitates

with carbonates, sulphide and hydroxide [68].

7. Zero-valent iron

ZVI tend to be oxidised, passing its electron to contaminants

(organic-halogenated hydrocarbons, inorganic- some metal;

U(VI), Cr(VI), etc., which undergo reductive mechanism

resulting in precipitation or degradation. The mechanism for

metal precipitation and degradation of halogenated hydro-

carbon by ZVI is given elsewhere [19,25,68]. In the case of

reductive precipitation, there is also a potential risk of

remobilization due to dissociation [33].

Interferences with reactions can occur while treating

groundwater with ZVI [20,68].

� O

xygen oxidises ZVI and causes FeO(OH) or Fe(OH)3 solids

to form. Besides making the ZVI no longer available to

interact with contaminants, the solid formation can clog the

reactive media, thus reducing the hydraulic permeability/

conductivity. In order to overcome this limitation, a pre-

treatment buffer zone of sand and pae gravel mixed with

about 10–15% ZVI by weight can be placed before the actual

treatment reactive barrier with 100% ZVI reactive cell. The

pre-treatment cell containing small amounts of ZVI will

remove the dissolved oxygen from the contaminated plume

Page 6: Permeable reactive barrier for groundwater remediation

R. Thiruvenkatachari et al. / Journal of Industrial and Engineering Chemistry 14 (2008) 145–156150

before it reaches the actual reactive cell. Also, most of the

precipitates formed in the pre-treatment stage will be retained

and as the permeability is higher, it will not affect the

groundwater flow to the main barrier [20].

� Z

VI interaction with the contaminant plume would also result

in the increase in pH, which results in the formation of

precipitates. The carbonate and hydroxide precipitates with

species like Fe, Ca, and Mg could impede the groundwater

flow through the barrier. Solid precipitation can be favored

when it results in the removal of toxic metals from the

groundwater. However, it is detrimental if it decreases the

hydraulic conductivity through the barrier or if it interferes

with the dominant contaminant removal mechanism. To

overcome this limitation, buffering amendments to prevent

the increase in pH is applied, as also reported earlier in this

report [20]. Alternative methods like ultrasound technology

to deal with permeability problem are also being evaluated

[33,69].

� U

nder anaerobic conditions, hydrogen gas is formed as a

product of iron corrosion which may also temporarily

passivate the iron surface [70]. Venting of H2 bubbles may be

required to maintain water flow and iron reactivity.

Alternatively, microorganism that can utilize hydrogen as

an energy source in anaerobic environment can also be

introduced [70,71].

� N

itrate has been shown to negatively impact reaction rates by

progressively passivating iron surface [72].

� E

arly indications suggest that high concentration of dissolved

silica also may have similar inhibitory effect on iron like

nitrates [73].

� C

ertain types of dissolved organic carbon (DOC) have been

shown to coat reactive sites on the iron, rendering it

unreactive [27].

Recent studies indicate that microorganisms with Fe0

increase the contaminant removal efficiency [74]. Application

of bio-augmentation in conjunction with the ZVI technology

and found that these two technologies can have a symbiotic

effect on each other [75–77]. Till et al. [75] identified that Fe0

can stoichiometrically reduce nitrate to ammonium and that

hydrogen produced (during anaerobic Fe0 corrosion by water)

can sustain microbial denitrification to reduce nitrate to more

innocuous products (i.e., N2O and N2).

Experiments with mixtures of contaminants have also shown

that bioaugmentation of PRBs with bacteria offers promise

when more than one contaminant is present. More complete

dechlorination occurred when the Fe0 was bioaugmented.

Batch experiments with mixtures of carbontetrachloride, Cr6+,

and nitrate showed that bioaugmentation reduced competition

by these pollutants for active sites on the Fe0 surface [74].

Permeable reactive barriers designed to enhance bacterial

sulfate reduction (sulfate reducing bacteria-SRB) and metal

sulfide precipitation have the potential to prevent acid mine

drainage and the associated release of dissolved metals. In this

situation, the conditions typically found within a reactive

barrier environment are well suited to SRB. Permeable reactive

barriers provide dissolved C, N, and P, and the plume water

entering the barrier provides high concentrations of iron and

other metals, provides the necessary condition and promotes

growth and reproduction of microorganisms [78]. A synergistic

interaction between microbial activity and ZVI obtained an

enhanced degradation efficiency of hydrocarbon pollutants,

according to the study by [79].

8. Activated carbon

Activated carbons are chemically stable materials and are

widely considered as suitable adsorbent for on-site or off-site

treatment of polluted groundwater [40]. This material presents

a high adsorption capacity for many organic and inorganic

contaminants largely due to its high surface area (about

1000 m2/g) and the presence of different types of surface

functional groups (hydroxyl, carbonyl, lactone, carboxylic

acid, etc.) [80]. In granular form, activated carbon appears to

be highly suitable for use in permeable barriers [34].

Significant removal of hexavalent chromium from contami-

nated groundwater using granular activated carbon (GAC) was

achieved by [34]. Regeneration of carbon by phosphate

extraction and acid washing also appeared to be successful

[34], allowing the possibility for repeated use of the material.

Microbial regeneration of activated carbon (used in organic

sorption) in PRB is a promising area, which needs to be

explored. A recent study by [81] with PRB using activated

carbon and microorganism for polyaromatic hydrocarbon

removal, showed that the degradation efficiency of organic

material was found to increase when the organic material was

adsorbed on the carbon. Few more studies on the organic

removal using activated carbon in PRBs have shown

promising outcomes [82]. Its effect on inorganic species is

yet to be evaluated in detail.

9. Zeolites

Zeolites are tectosilicates with three-dimensional alumino-

silicates structure containing water molecules, alkali and

alkaline earth metals in their structural framework [83]. These

minerals have very high ion-exchange, adsorbing, catalytic,

molecular sieving capacities and make them potentially useful

as treatment mineral for use in the PRBs [18,27]. As the mineral

is anionic (negatively charged), it can be used to remove cations

from aqueous solutions. Several hundred zeolitic materials

exists; Clinoptilolite, a natural zeolite, is a potential material for

remediation of aqueous solutions since it demonstrates strong

affinity for several toxic heavy metals [84] and can selectively

adsorb some radionuclides [85]. A new surfactant modified

zeolite has been tested to simultaneously remove organic and

inorganic species [19].

10. Alkaline materials-complexing agents

Hydrated lime (Ca(OH)2) is a cheap reagent which can be

used in PRBs for groundwater remediation. Previously, it has

been used for remediation of acid mine drainage [18,27].

Lime barriers cause pH to increase to 12–12.5 in order to

Page 7: Permeable reactive barrier for groundwater remediation

Table 4

Advantages and limitations of various reactive barriers used in PRB technology

Material Advantages Limitations

Zero valent iron (ZVI) Most commonly used reactive barrier material.

Vast amount of background data available

Not effective on all types of organic compounds especially

certain dense non-aqueous phase liquid compounds like

1,2-dichloroethane and dichloromethane [54]

Ability to be used in different states: as a

pile [24,90], powder/granular [62],

filing [91], colloidal [58,59], nanosized [61,62],

emulsion that can be injected [63]

Lifetime of the material could be reduced due to the

formation of surface coating due to geological condition

of the site [93]

High reactivity with organic and inorganic

contaminants [92]

Increase in pH during reaction induces corrosion and

subsequent precipitation of minerals would lead to

decreased permeability of reactive material [94]

Ability to combine with other treatment

methods, e.g. bioremediation

H2 gas produced and the microorganism (biofouling)

could reduce the porosity of reactive material

Less or no major problems associated with

occupational health and safety (OHS) in

handling this material

Limited information available on long term performance

of the system especially on the build-up of surface

precipitates and biofouling.

Compounds like silica or natural organic matter (NOM)

have passivating effect; reducing the iron reactivity [18,93]

Competitive reaction inhibits the reactivity in the presence of

certain compounds. In the presence of nitrate the

dehalogenation of chlorinated compounds is decreased

[95,96]

Activated carbon Different types (with different reactivities) of

activated carbon can be obtained from low cost

natural products, e.g. coconut shell based

Vast data on ex situ water and wastewater treatment.

But very limited data on in situ treatment under field

conditions

Effective in the treatment of organic and heavy

metal contaminants [97]

Rapid breakthrough and thus frequent carbon change-outs

or regeneration. Requires optimization studies

Excellent material to combine with biotreatment Performance highly dependent on temperature and other

extrinsic parameters

Chemically stable material Surface coatings may decrease sorption capacity

Competitive adsorption

Lime (calcium carbonate

or hydroxide)

Low cost reactive PRB material Slow reaction time

Effective in neutralization; reducing the solubility of

certain metals or conditioning hydrochemical

system to assist with other treatment processes,

e.g. bioremediation

Loss in efficiency of the system because of coating of

the limestone particles with iron precipitates

Used extensively for acid mine drainage remediation or

acidic agricultural soils

Difficulty in treating acid mine drainage with a high

ferrous–ferric ratio, and ineffectiveness in removing

manganese

Limestone treatment is generally not effective for acidities

exceeding 50 mg/L

Voluminous sludge is produced with hydrated lime

(calcium hydroxide)

Microbial bioremediation Less expensive and easy to install A perceived lack of knowledge about biodegradation

mechanism

Natural processes to treat contaminants Specific contaminants may not be amenable to

biodegradation [98]

Capability to degrade organic contaminants into relatively

less toxic end products

In the case of mixed wastes, some are amenable only under

aerobic condition and some only under aerobic condition

Reduced risk of human exposure to contaminated media Site characterization and optimization studies are required for

each contaminated site. Chemical characteristics of the

contaminants dictate the extent of biodegradability

Remediation is not restricted to the treatment zone alone

Works well for dissolved contaminants and contamination

adsorbed onto higher permeability sediments

(sands and gravels)

R. Thiruvenkatachari et al. / Journal of Industrial and Engineering Chemistry 14 (2008) 145–156 151

facilitate the formation of metal hydroxides, which reduce the

solubility of certain metals. They have proved to be

successful in remediation of anionic and cationic pollutant

species [27].

11. Bioremediation

Remediation of pollutants using microorganisms (bioreme-

diation) is one of the promising and viable technologies, which

Page 8: Permeable reactive barrier for groundwater remediation

Table 5

Examples of sequential PRBs

� Granular zero-valent iron to treat chlorinated hydrocarbons followed by

aerobic bioremediation (using oxygen release compounds) to treat aromatic

hydrocarbons

� Granular zero-valent iron to treat chlorinated hydrocarbons followed by

nutrient addition or solid carbon source addition to promote anaerobic

biodegradation of volatile organic carbons (VOCs) that cannot be degraded by

granular iron

� Sequential anaerobic biodegradation of chlorinated solvents followed by

aerobic biodegradation for toluene degradation

� Solid carbon sources to treat nitrate followed by granular iron to treat VOCs

� A four component PRB consisting of four reactive media in series for treating

(immobilizing or destroying) multiple contaminants

R. Thiruvenkatachari et al. / Journal of Industrial and Engineering Chemistry 14 (2008) 145–156152

utilizes naturally occurring microorganisms and biological

reactions for the degradation of toxic contaminants (mostly

organic and in some instances inorganic compounds) in the

subsurface groundwater. In situ bioremediation creates subsur-

face environmental conditions (without withdrawing the

contaminated water from under the ground), typically through

the principle of oxidation–reduction manipulation, which

induce the degradation of chemicals via microbial catalyzed

biochemical reactions [45].

Many environmental pollutants such as petroleum hydro-

carbons are highly reduced, which allows oxidation of these

pollutants to innocuous final compounds like carbon dioxide

and water. On the other hand some pollutants like chlorinated

solvents and nitrates are highly oxidized and are prone to

undergo reduction. Microorganisms mediate such redox

reactions (anaerobically or aerobically) and thrive on the

contaminant degradation process to obtain it energy and food

for its growth [33].

11.1. Organic carbon for denitrification and sulphate

reduction

Organic carbon for denitrification and sulfate reduction has

been well recognized. Robertson and Cherry [86] adapted the

use of permeable organic carbon material to stimulate

biologically mediated denitrification and sulfate reduction in

contaminated groundwater in PRB system. Denitritrifying and

sulfate reducing bacteria are ubiquitous in the environment. In

the presence of organic source, these heterotrophic bacteria

reduce nitrate to nitrogen gas and sulphate to sulphide, in the

absence of oxygen [35,87–89]. Generation of sulphide during

sulphate reduction, precipitate the soluble metals or metalloids

as low-solubility sulphide minerals. In combination, the

sulphate reduction and sulphide precipitation reaction have

the potential effect of decreasing concentrations of sulphate,

iron, and other metals and metalloids and increasing alkalinity

and pH. Table 4 shows the advantages and limitations of various

reactive barriers used in PRB technology.

12. Sequential reactive media

Two of the most common classes of organic contaminants in

groundwater are chlorinated solvents and petroleum derived

aromatics. These chemicals are known to contaminate vast

quantities of water and are difficult to remediate. Chlorinated

solvents including chlorinated ethenes and chloromethanes

have densities grater than water, allowing them to penetrate far

below the water table. In addition, they have very low

solubilities in water and are resistant to degradation. Even if

they do degrade, it is often along pathways that can lead to

products of greater concern than the original contaminants.

These characteristics have led to the phrase ‘‘dense nonaqueous

phase liquids’’ (DNAPL) to describe such chemicals [99].

The chlorinated solvents contain carbon in a relatively

oxidized form, while, in contrast, the petroleum derived

organics contain carbon in a relatively reduced form. Hence the

geochemical conditions required for degradation of these

classes of chemical in an aquifer are quite different. Often

mixtures of such contaminants are encountered and in many

occasions PRBs with single media would be unable to deal with

such situations. This has led to the development of new

strategies for the cleanup of contaminated aquifer.

The PRB with sequential treatment with duel or more kinds

of reactive barriers have recently been given more attention

(Fiorenza et al., 2000). Some of the sequential PRBs adopted

are given in Table 5 [27,100,101].

13. In situ chemical oxidation (ISCO)

ISCO can be applied to a variety of soil types and sizes and

can treat volatile organic chemicals (VOCs) including

dichloroethene (DCE), trichloroethene (TCE), tetrachlor-

oethene (PCE), and benzene, toluene, ethylbenzene, and

xylene (BTEX) as well as semi-volatile organic chemicals

(SVOCs) including pesticides, polycyclic aromatic hydrocar-

bons (PAHs), and polychlorinated biphenyls (PCBs). The

oxidants such as hydrogen peroxide (H2O2), potassium

permanganete (KMnO4), ozone (O3) are applied [102,103].

The mechanism of degradation is the production of hydroxyl

radicals, which are capable of oxidizing complex organic

compounds. For most in situ reactive zone targets, hydroxyl

radical oxidation is a much faster treatment method [104].

Some of the advantages and limitations in adopting ISCO is

given in Table 6.

13.1. Use of oxidants

(i) Potassium or sodium permanganate (KMnO4/NaMnO4):

Permanaganate is an oxidizing agent and has an affinity

towards organic compounds containing carbon-carbon

bonds, aldehyde groups, or hydroxyl groups [105–107]. As

the permanganate dose increases, organic degradation also

increases and the permanganate consumption also

increases [27]. In some instances where permanganate is

used for the degradation of DNAPL, production of MnO2

(s) may significantly lower the permeability of the soil

matrix and form a coating on DNAPL [27,108]. Phase

transfer catalyst (like quaternary amines), which has both

polar and non-polar segments, are also used along with

permanganate, to assist in the removal hydrophobic

Page 9: Permeable reactive barrier for groundwater remediation

Table 6

Advantages and limitations in using ISCO method in PRB technology

Advantages Limitations

Rapid treatment Limited information is available on the operational history

Capability of contaminants to be degraded into final

innocuous compounds like carbon dioxide, water and chloride

Proper oxidant selection must be made depending on site

characteristics and nature of contaminants

Oxidants can be injected and can be applied to greater depths

compared to method which require excavation

Problems such as gas evolution (also explosive vapors), toxic by-products,

resolubilization of metals and reduction of biomass

Also suitable for complex organics (recalcitrant compounds)

and mixed waste streams

Precise control of pH, temperature, and contact time is important

Certain materials like natural organic matter (NOM), reduced inorganics such

as iron and manganese can exert a demand for oxidant

Occupational health and safety concerns

Table

Char

Techn

Physi

Key

Oxid

By-p

Reag

Subsu

R. Thiruvenkatachari et al. / Journal of Industrial and Engineering Chemistry 14 (2008) 145–156 153

organic contaminants [104]. Permanganate reaction rate is

much slower than those of hydroxyl radical reactions,

which gives it s significant advantage in achieving greater

contact with contaminats, especially with NADPL

compounds [104].

(ii) H

ydrogen peroxide (H2O2)/fenton (H2O2–Fe) oxidation:

Concentration of peroxide injection generally varies

between 3% and 35% (w/w) at a pH range of 3.5–6, with

or without a metal catalyst (usually iron). The process

involves the production of hydroxyl radicals (OH.), which

act as a powerful oxidizing agent for the degradation of

organic contaminants. The initial amount of H2O2 and Fe

catalyst also depends on the contaminant level in the soil,

and the volume of groundwater remediated [27].

Hydrogen peroxide and fenton oxidation has been

applied for the remediation of several contaminants

including petroleum hydrocarbons [109,110], hexadecane

[111], atrazine [112,113], aromatic compounds [114,115],

dioxins [116] and chlorinated solvents [117]. The

usefulness of fenton oxidation may be limited by low

soil permeability, subsurface heterogeneities, and highly

alkaline soil (where carbonate ions are free radical

scavengers) [105].

(iii) P

ersulfate: Sodium persulfate (Na2S2O8) is recently being

used in ISCO. The conversion of the persulfate results in

the production of sulfate radicals or a single electron

radical. Free radicals act as strong oxidizing agents and are

known to oxidize many VOCs. Many sites have sufficient

background conditions which allow the persulfate to

oxidize the targeted VOCs present in the soils and

groundwater without the addition of a catalyst. In some

7

acteristics of common oxidants

ology features Fenton’s reagent Permanganate

cal state as injected Liquid Liquid

oxidant OH� MnO4�

ation potential 2.8 V 1.7 V

roducts Fe(III), O2, H2O Mn(VI)

ent costs Moderate Moderate for KMn

rface fouling Possible Yes, due to MnO2

instances, chelated iron compounds have been used as

activators [27,118].

(iv) O

zone (O3): Ozone is a very powerful oxidant and is

commonly used for the remediation of hydrocarbon and

chlorinated solvent contaminants [27,119,120]. Ozone

solubility and the concentration of ozone in the injected

gas stream are the key variables affecting the rate of

treatment. Ozone degrades to oxygen and promotes

biodegradation in combination to ozone oxidation process

[27,121]. Treatment with ozone in combination of

hydrogen peroxide (termed as peroxone process) [104]

and fenton reagents [122] has also been attempted.

The characteristics of various oxidants used in the ISCO

technology are given in Table 7.

14. PRB studies in Australia and New Zealand

� Several sightings on petroleum hydrocarbon contamination

have been reported.

- BTEX (benzene, toluene, ethylbenzene and xylene)

contamination in metropolitan Perth, Western Australia.

Biodegradation has been adopted as the remedial method.

Preferential intrinsic biodegradation of selected organic

compounds within the BTEX plume was shown to be

occurring [98].

- BTEX and other dissolved organic contamination from

gasoline spills at Kwinana field on the Swan Coastal Plain

in Western Australia. Aerobic bioremediation was adopted.

Volatalization found to be a dominant mechanism of

removal of the VOCs contaminant present. Evidence of

Persulfate Ozone

Liquid Gas

SO42� O3 and OH�

2.5 V 2.07 V/2.8 V

Sulfate Oxygen

O4; high for NaMnO4 Moderate Moderate

formation No No

Page 10: Permeable reactive barrier for groundwater remediation

R. Thiruvenkatachari et al. / Journal of Industrial and Engineering Chemistry 14 (2008) 145–156154

biodegradation of dissolved organics was also identified

from the oxygen utilization rates [123].

- Trace concentrations of petroleum hydrocarbon–nonaro-

matic hydrocarbons (BTEX) and chlorinated solvents-

halogenated aliphatic hydrocarbon (tetrachloroethene-

PCE) at the Swan Coastal Plain area, Western Australia

was treated by biodegradation method. The results from the

column studies indicated that indigenous microorganisms

were capable of degrading trace amounts of toluene and

ethylbenzene under denitrifying conditions and toluene

under sulfate-reducing conditions. Benzene was persistent

under anoxic conditions, but degraded readily under

oxygenated condition. PCE failed to degrade under either

oxic or anoxic conditions.

- A funnel and gate PRB was adopted with saturated peat

(70% biogreen humic reed sedge, 20% sphagnum peat and

10% cocoa fibre) reactive material was tested for the

remediation of toluene, ethylbenzene, xylene and n-alkane

contaminants (spirit petroleum hydrocarbon spill at the

underground storage site in South Eastern Australia). This

study [124] reported high sorptive efficiencies compared to

previous study which shows a decrease in removal

performance with time [125]. This reinforces the impor-

tance of testing local peat types prior to full-scale

application.

� L

eachate contamination: Coal washery alkaline slag lea-

chate, Kemblawarra, Port Kembla, NSW, Australia [68]. The

wastes include air-cooled blast furnace slag, steel making

slag and coal washery discard. A PRB wall with coal washery

discard was chosen. The geochemical calculations indicate

that the reactive wall’s ability to control alkalinity and S2�

should outlast the slag’s ability to produce toxic leachate

[68,126].

� D

NAPL contamination: A pilot scale zero valent iron reactive

barrier was tested for the remediation of DNAPL, including

trichloroethene (TCE), tetrachloroethene (PCE), carbon

tetrachloride (CTC), 1,1,2,2-tetrachloroetane (PCA), 1,1,1-

trichloroethane, 1,2-dichloroethane, chloroform, chloroben-

zene, hexachlorobutadiene and hexachloroethane, at South of

City of Sydney, Australia [127]. Very high removal

efficiencies of VOC was observed. However, very low

removal performance was noted for compounds such as 1,2-

dichloroethane and dichloromethane [54].

� A

mmonium contaminated groundwater: Pilot scale, sequen-

tial use of polymer mat reactive barrier for the remediation of

ammonium-contaminated groundwater [128]. Two polymer

mats were placed in series along the flow of the contaminant

plume. The upgradient mat delivered oxygen to induce

bacterial nitrification of the ammonium to nitrite/nitrate and

the downgradient mat delivered ethanol to induce bacterial

denitrification of the nitrite/nitrate to produce nitrogen gas.

Significant reduction (>90%) in total N was achieved.

� D

enitrification (Bardowie farm, Cambridge, North Island,

New Zealand): Pilot scale denitrification wall was con-

structed by mixing a carbon source, such as sawdust (Pinus

radiata) (40 m3), in the subsurface aquifer. The added carbon

source stimulates nitrate reduction to nitrogen gas by

denitrification. However, it was found that carbon availability

was not limiting the size of the microbial population [129].

� A

cid rock drainage (AMD) treatment: Remediation of acid

mine drainage was attempted at Mt Carrington silver and gold

mine, in northern New South Wales, Australia. BauxsolTM, a

product made from seawater-neutralized red mud (a by-

product of alumina refining) was used as the reactive barrier

material. This material has very low hydraulic conductivity

(as it is very fine-grained (<10 mm)), it was mixed with sand.

Over 45,000 L of ARD was treated above ground with

Bauxsol/sand PRB material. It was found that Bauxsol/sand

PRB was able to avoid clogging (which is a major problem

when using carbonates and hydroxides for acid neutraliza-

tion) and achieve reductions satisfactory with Australian

Water quality Standards. Metal removal efficiency was also

very high. Analysis of the spent Bauxsol also indicated that

the bound metals are not readily leachable, allowing the spent

material to be disposed safely in landfills.

� B

iodegradation of pesticides: Contamination of groundwater

by atrazine (2-chloro-4-ethylamino-6-isopropylamino-1,3,5-

triazine), terbutryn (2-tert-butylamino-4-ethylamino-6-

methylthio-1,3,5-triazine) and fenamiphos (1-(methy-

lethyl)-ethyl-3-methyl-4-(methylthio) phenylphosphorami-

date) from infiltration of pesticide-laden washwater from

Dianella, Perth, Western Australia, was treated using a lab-

scale column experiments by biodegradation [130]. Polymer

mats were used to deliver oxygen into the groundwater to

induce biodegradation. Degradation rates of atrizine were

relatively high, however, no significant degradation of

terbutryn or fenamiphos was observed.

Acknowledgement

This study was funded by the CRC Care (Land and Water),

Australia. At the time of the study the first author was with the

University of Technology Sydney.

References

[1] UK Environmental Agency, http://www.environment-agency.gov.uk/,

2005.

[2] S.S.D. Foster, in: O. Sililo (Ed.), Groundwater-Past Achievements and

Future Challenges, Balkema, Rotterdam, The Netherlands, 2000, p. 27.

[3] UNEP, Groundwater and its susceptibility to degradation: a global

assessment of the problem and options for management, early warning

and assessment report series, UNEP/DEWA/RS. 03-3, joint publication

from United Nations Environment Programme, Department for Interna-

tional Development and the Natural Environment Research Council

(NERC), 2003.

[4] Land and Water Australia, Australian Water Resource Assessment 2000

– Surface Water and Groundwater – Availability and Quality, Land and

Water Resource Audit, Commonwealth of Australia, 2001.

[5] Australia State of Environment Report, Inland Waters, Commonwealth

of Australia, 2001.

[6] Water and Rivers Commission, Water Facts, Department of Western

Australia, 1998.

[7] Ministry of Land and Water Conservation, The NSW State Government

Policy Framework Document, NSW Government, Australia, 1997.

[8] P.B. Bedient, H.S. Rifai, C.J. Newell, Groundwater Contamination:

Transport and Remediation, PTR Prentice Hall, Inc., USA, 1994.

Page 11: Permeable reactive barrier for groundwater remediation

R. Thiruvenkatachari et al. / Journal of Industrial and Engineering Chemistry 14 (2008) 145–156 155

[9] USEPA, Evaluation of Groundwater Extraction Remedies, vols. 1 and 2,

EPA Office of Emergency and Remedial Responses, Washington, DC,

1989.

[10] USEPA, Evaluation of Groundwater Extraction Remedies: Phase II, vol.

1, Summary Report, Publication 9355.4-05, EPA Office of Emergency

and Remedial Responses, Washington, DC, 1992.

[11] National Research Council (NRC), Alternatives for Groundwater

Cleanup, National Academy Press, Washington, DC, 1994.

[12] NRC, Contaminants in the Subsurface: Source Zone Assessment and

Remediation, National Academy Press, Washington, DC, USA, 2005.

[13] J. Ball, Inland Waters-Theme Report, Australia State of the Environment

Report 2001 (Theme Report), CSIRO Publishing on behalf of the

Department of the Environment and Heritage, Commonwealth of Aus-

tralia, Website: http://www.deh.gov.au/soe/2001/inland/, 2001.

[14] SKM, Impact of Logging Practices on Water Yield and Quality in the

Otway Forests, Department of Natural Resources and Environment,

Australia, 2000.

[15] LWRRDC, Contamination of Australian Groundwater Systems with

Nitrate, Occasional Paper No. 03/99, Land and Water Resources

Research and Development Corporation, Canberra, 1999.

[16] NHMRC/ARMCANZ, Australian Drinking Water Guidelines, National

Health and Medical Research Council & Agricultural and Resource

Management, Council of Australia and New Zealand, Canberra, 1996.

[17] N.J. Schofield, B.W. Simpson, Aust. J. Water Resour. 1 (2) (1996) 91.

[18] K.E. Roehl, T. Meggyes, F.G. Simon, D.I. Stewart, Long-term Perfor-

mance of Permeable Reactive Barriers, Elsevier Publishers, 2005.

[19] A. Xenidis, A. Moirou, I. Paspaliaris, J. Miner. Wealth 123 (2002) 35.

[20] A.R. Gavaskar, N. Gupta, B.M. Sass, R.J. Janosy, D. O’Sullivan, Perme-

able Barriers for Groundwater Remediation: Design, Construction, and

Monitoring, Battelle Press, Columbus, Ohio, 1998.

[21] R.D. Vidic, Permeable Reactive Barriers: Case Study Review, Technol-

ogy Evaluation Report TE01-01, Groundwater Remediation Technolo-

gies Analysis Center, 2001.

[22] R.D. Vidic, F.G. Pohland, Treatment wells, Technology Evaluation

Report TE 96-01. GWRTAC, Pittsburg, USA, 1996.

[24] A.R. Gavaskar, J. Hazard. Mater. 68 (1999) 41.

[25] T. Meggyes, F.-G. Simon, J. Land Contam. Remediat. 8 (3) (2000) 1.

[26] Golder Associates Ltd., Active Containment: Combined Treatment and

Contaminant Systems, Department of Environment, Transport and the

Regions, ISBN 1851121145, London, 1998.

[27] ITRC, Permeable Reactive Barriers: Lessons Learned/New Directions.

Technical/Regulatory Guidelines, ITRC, Washington, DC, 2005.

[29] F.-G. Simon, T. Tunnermeier, T. Meggyes, in: G. Prokop (Ed.), Proceed-

ings 1st IMAGE-TRAIN Cluster Meeting, Karlsruhe, November 7–9,

2001, (2002), p. 145, Vienna (Umweltbundesamt).

[30] PEREBAR, 5th Framework Programme Research and Technology

Development Project on Long-term Performance of Permeable Reactive

Barriers Used for the Remediation of Contaminated Groundwater.

PEREBAR EVK1-CT-1999-00035, National Technical University of

Athens, Greece, 2002.

[31] USEPA, Permeable reactive barrier technologies for contaminant reme-

diation, Report number EPA/600/R-98/125, Washington, DC, 1998.

[32] P.G. Tratnyek, M.M. Scherer, T.L. Johnson, L.J. Matheson, in: M.A. Tarr

(Ed.), Chemical Degradation Methods for Wastes and Pollutants: Envir-

onmental and Industrial Application, Marcel Dekker, New York, 2003, p.

371.

[33] M.M. Scherer, S. Richter, R.L. Valentine, P.J.J. Alvarez, Environ. Sci.

Technol. 30 (3) (2000) 363.

[34] I. Han, M.A. Schlautman, B. Batchelor, Water Environ. Res. 72 (1)

(2000) 29.

[35] S.G. Benner, D.W. Blowes, W.D. Gould, R.B. Herbert, C.J. Ptacek,

Environ. Sci. Technol. 33 (1999) 2793.

[36] R.S. Hedin, G.R. Watlaf, in: Proceedings, International Land Reclama-

tion and Mine Drainage Conference and Third International Conference

on the Abatement of Acid Drainage, 1, Pittsburg, Pensylvania, (1994), p.

184.

[37] W. Admassu, T. Breese, J. Hazard. Mater. B 69 (1999) 187.

[38] M. Fuhrman, D. Aloysius, H. Zhou, J. Waste Manage. 15 (7) (1995) 485.

[39] P. Misaelides, A. Godelitsas, D. Filippidis, D. Charistos, I. Anousis, Sci.

Total Environ. 173/174 (1995) 237.

[40] P. Huttenlock, K.E. Roehl, K. Czurda, Environ. Sci. Technol. 35 (2001)

4260.

[41] C.L. Ake, K. Mayura, H. Huebner, G.R. Bratton, T.D. Phillips, J. Toxicol.

Environ. Health, Part A 63 (2001) 459.

[42] M.J. Baker, D.W. Blowes, C.J. Ptacek, Environ. Sci. Technol. 32 (1998)

2308.

[43] R. Wilson, D.M. Mackay, K.M. Scow, Environ. Sci. Technol. 36 (2)

(2002) 190.

[44] A. Abdelouas, W. Lutze, W. Gong, E.H. Nuttall, B.A. Strietelmeier, B.J.

Travis, Sci. Total Envrion. 250 (1–3) (2000) 21.

[45] ITRC, A Systematic Approach to In situ Bioremediation in Groundwater,

Including Decision Tree on In situ Bioremediation for Nitrates, Carbon

Tetrachloride, and Perchlorate, ITRC, USA, 2002.

[46] USEPA, Engineered Approaches to In Situ Bioremediation of Chlori-

nated Solvents: Fundamentals and Field Applications. EPA 542-R-00-

008, Office of Solid Waste and Emergency Response, Technology

Innovation Office, USEPA, Washington, DC, 2000.

[47] F. Caruso, D. Trau, H. Mohwald, R. Renneberg, Langmuir 16 (2000)

1485.

[48] PEREBAR, 5th Framework Programme Research and Technology Pro-

ject on Long-term Performance of Permeable Reactive Barriers Used for

the Remediation of Contaminated Groundwater, University of Leeds,

UK, 2003.

[49] T. Shimotori, E. Nuxoll, E. Cussler, W.A. Arnold, Environ. Sci. Technol.

38 (2004) 2264.

[50] J.A. Sacre, Treatment Walls: A Status Update, Groundwater Remedia-

tion Technologies Analysis Center, TP-97-02, Pitsburg, PA, Homepage:

http://www.gwrtac.org, 1997.

[51] USEPA, Fiels Applications of In situ Remediation Technologies: Perme-

able Reactive Barriers, Office of Solid Waste and Emergency Response

Technology Innovation Office, Washington, DC, 2002.

[52] A. Agarwal, P.G. Tratnyek, Environ. Sci. Technol. 30 (1) (1996) 153.

[53] L.J. Matheson, P.G. Tratnyek, Environ. Sci. Technol. 28 (1994) 2045.

[54] R.W. Gillham, S.F. O’Hannesin, Ground Water 32 (6) (1994) 958.

[55] T.L. Johnson, M.M. Scherer, P.G. Tratnyek, Environ. Sci. Technol. 30 (8)

(1996) 2634.

[56] D.R. Burris, T.J. Campbell, V.S. Manoranjan, Environ. Sci. Technol. 29

(11) (1995) 2850.

[57] R.A. Holser, S.C. McCutcheon, N.L. Wolfe, in: Mass Transfer Effect on

the Dehalogenation of Trichloroethene by Iron/pyrite Mixtures.

Extended Abstracts from the 209th ACS National Meeting, Anaheim,

CA 35(1) p. 788, Division of Environmental Chemistry, American

Chemical Society, Washington, DC, 1995.

[58] K.J. Cantrell, D.I. Kaplan, T.J. Gilmore, J. Environ. Eng. 123 (8) (1997)

786.

[59] D.I. Kaplan, K.J. Cantrell, T.W. Wietsma, M.A. Potter, J. Environ. Qual.

25 (1996) 1086.

[60] H.-L. Lien, W.-X. Zhang, J. Environ. Eng. 125 (11) (1999) 1042.

[61] S.H. Joo, et al. Environ. Sci. Technol. 38 (2004) 2242.

[62] C.-B. Wang, W.-X. Zhang, Environ. Sci. Technol. 31 (7) (1997) 2154.

[63] J. Quinn, et al. Environ. Sci. Technol. 39 (2005) 1309.

[64] R. Muftikian, Q. Fernando, N. Korte, Water Res. 29 (1995) 2434.

[65] N.E. Korte, L. Liang, J. Clausen, Emerging Technologies in Hazardous

Waste Management VII, Extended Abstract for the Special Symposium,

Atlanta, GA, (1995), p. 45.

[66] E.L. Appleton, Environ. Sci. Technol. 30 (12) (1996) 536A.

[67] T.M. Sivavec, D.P. Horney, in: Reductive Dechlorination of Chlorinated

Ethenes by Iron Metal. Presented at the 209th ACS National Meeting,

Anaheim, CA, April 2–6, 1995.

[68] N. Ott, Permeable Reactive Barriers for Inorganics, USEPA, Washington,

DC, Report obtained from website http://www.clu-in.org, 2000.

[69] C.L. Geiger, C.A. Clausen, D.R. Reinhart, N. Ruiz, K. Farrell, A.S.

Sonawane, Using Ultrasound for Restoring Iron Activity in Permeable

Treatment Walls. 221st National Meeting, American Chemical Society.

San Diego, CA, 2001, p. 1167 (Preprint Extended Abstracts, Division of

Environmental Chemistry 41, No. 1).

Page 12: Permeable reactive barrier for groundwater remediation

R. Thiruvenkatachari et al. / Journal of Industrial and Engineering Chemistry 14 (2008) 145–156156

[70] R.T. Wilkin, R.W. Puls, G.W. Sewell, Long-term Performance of Perme-

able Reactive Barriers Using Zero-valent Iron: An evaluation at Two

Sites, Environmental Research Brief, EPA/600/S-02/001, 2002.

[71] R.M. Atlas, R. Bartha, Microbial Ecology, third ed., Benjamin Cum-

mings, New York, 1993.

[72] K. Ritter, M.S. Odziemkowski, R.W. Gillham, J. Contam. Hydrol. 55

(2002) 87.

[73] T. Kohn, S.R. Kane, D.H. Fairbrother, A.L. Roberts, Environ. Sci.

Technol. 37 (24) (2003) 5806.

[74] G.F. Parkin, P.J. Alvarez, M.M. Scherer, J.L. Schnoor, Role of Microbes

in Remediation with Fe0 Reactive Barriers, Ground Water Currents

USEPA 542-N-00-002 issue no. 35, p. 2, website http://www.clu-in.

org/products/newsltrs/gwc/gwc0300.htm, 2000.

[75] B.A. Till, L.J. Weathers, P.J.J. Alvarez, Environ. Sci. Technol. 32 (5)

(1998) 634.

[76] ARS Technologies Inc., http://www.arstechnologies.com/, Website vis-

ited January 2005, 2005.

[77] B.-T. Oh, C.L. Just, P.J. Alvarez, Environ. Sci. Technol. 35 (2001) 4341.

[78] K.R. Waybrant, C.J. Ptacek, D.W. Blowes, Environ. Sci. Technol. 36

(2002) 1349.

[79] B. Wrenn, Enhanced Reductive Dechlorination through Biological Inter-

action with Zero-valent Iron, Federal Remediation Technologies Round-

table Meeting, Arlington, Virginia, USA, Website: http://www.frtr.gov/

pdf/meetings/summary_09jun04.pdf, 2004.

[80] M.O. Corapcioglu, C.P. Huang, Carbons 25 (1987) 569.

[81] P. Leglize, Bacteria-phenanthrene-activated Carbon Interactions Studies

for PRB Process Evaluation, Dissertaion, Faculty of Sciences and

Technology of Nancy, France, ADEME, Research Programme Planning,

Cedex, France, Website: http://www.ademe.fr/anglais/thesis/thesis04/

vath201204.htm, 2004.

[82] H. Schad, P. Gratwohl, in: H. Burmier (Ed.), Treatment Walls and

Permeable Reactive Barriers, vol. 229, pp. 56065, NATO CCMS, Vienna,

1998.

[83] G. Gottardi, E. Galli, Natural Zeolites, Springer, Berlin, 1985, p. 409.

[84] M. Loizidou, R.P. Townsend, Zeolites 7 (2) (1987) 153.

[85] D. Lepperd, J. Mining Eng. (1990) 604 (June Issue).

[86] W.D. Robertson, J.A. Cherry, Ground Water 33 (1995) 99.

[87] W.J. Hunter, R.F. Follett, J.W. Cary, Am. Soc. Agric. Eng. 40 (1997) 345.

[88] S.G. Benner, D.W. Blowes, C.J. Ptacek, Groundwater Monit. Remediat.

17 (4) (1997) 99.

[89] W.D. Robertson, D.W. Blowes, C.J. Ptacek, J.A. Cherry, Ground Water

38 (5) (2000) 99.

[90] R.D. Vidic, F.G. Pohland, Treatment Wells, Technology Evaluation

Report TE 96-01, GWRTAC, Pittsburg, USA, 1996.

[91] A.M. Moore, C.H. De Leon, T.M. Young, Environ. Sci. Technol. 37 (14)

(2003) 3189.

[92] D.W. Blowes, C.J. Ptacek, S.G. Benner, C.W.T. McRae, T.A. Bennett,

R.W. Puls, J. Contam. Hydrol. 45 (2000) 123.

[93] J. Klausen, P.J. Vikesland, T. Kohn, D.R. Burris, W.P. Ball, A.L. Roberts,

Environ. Sci. Technol. 37 (2003) 1208.

[94] D.H. Phillips, B. Gu, D.B. Watson, Y. Roh, L. Liang, S.Y. Lee, Environ.

Sci. Technol. 34 (2000) 4169.

[95] J. Farrell, M. Kason, N. Melitas, T. Li, Environ. Sci. Technol. 34 (2000)

514.

[96] O. Schlicker, M. Ebert, M. Fruth, M. Weidner, W. Wust, A. Dahmke,

Ground Water 38 (3) (2000) 403.

[97] I. Snape, C.E. Morris, C.M. Cole, Cold Reg. Sci. Technol. 32 (2–3)

(2001) 157.

[98] G.B. Davis, C. Barber, T.R. Power, J. Thierrin, B.M. Patterson, J.L.

Rayner, Q. Wu, J. Contam. Hydrol. 36 (1999) 265.

[99] ITRC, Dense Non-aqueous Phase Liquids (DNAPLs): Review of Emer-

ging Characterization and Remedial Technologies. Technology Over-

view, ITRC Work Group DNAPLs/Chemical Oxidation Work Team,

USA, 2000.

[100] S. Fiorenza, C.L. Oubre, C.H. Ward, Sequenced Reactive Barriers for

Groundwater Remediation, Lewis Publishers, New York, NY, 2000.

[101] J.P. Kaszuba, P. Longmire, B.A. Strietelmeier, T.P. Taylor, D. Counce,

P.S. den Baars, T. Cota, J. Myers, R.S. Johnson, Demonstration of a

Multi-Layered Permeable Reactive Barrier in Mortandad Canyon at Los

Alamos National Laboratory. LA-UR-03-7320. Los Alamos, N.M., Los

Alamos National Laboratory, 2003.

[102] Environmental Security Technology Certification Program (ESTCP),

Technology Status Review In situ Oxidation, USA, 1999.

[103] USEPA, A Citizen’s Guide to Chemical Oxidation, EPA 542-F-01-013,

April 2001, Washington, DC, 2001.

[104] S.S. Suthersan, F.C. Payne, In situ Remediation Engineering, CRC Press,

USA, 2005.

[105] ITRC, Technical and Regulatory In situ Chemical Oxidation of Con-

taminated Soil and Groundwater, Technical/Regulatory Guidelines,

ITRC Work Group and In situ Chemical Oxidation Work Team, USA,

2001.

[106] B.S. Tunnicliffe, N.R. Thomson, J. Contam. Hydrol. 75 (2004) 91.

[107] L.K. Mackinnon, N.R. Thomson, J. Contam. Hydrol. 56 (2002) 49.

[108] M.H. Schroth, M. Oostrom, T.W. Wietsma, J.D. Istok, J. Contam. Hydrol.

50 (2001) 79.

[109] B.V. Moran, P. Babaian, C.P. Young, Annual Conference on Contami-

nated Soils, University of Massachusetts, 1998.

[110] R.J. Watts, D.R. Haller, A.P. Jones, A.L. Teel, J. Hazard. Mater. 76 (1)

(2000) 73.

[111] R.J. Watts, P.C. Stanton, Water Res. 33 (6) (1999) 1405.

[112] S.M. Arnold, W.J. Hickey, R.F. Harris, Environ. Sci. Technol. 29 (8)

(1995) 2083.

[113] R. Mecozzi, L.D. Palma, C. Merli, Chemosphere 62 (9) (2006)

1481.

[114] I. Casero, D. Sicilia, S. Rubio, D. Perez-Bendito, Water Res. 31 (8)

(1997) 1985.

[115] L. Lunar, D. Sicilia, S. Rubio, D. Perez-Bendito, U. Nickel, Water Res.

34 (6) (2000) 1791.

[116] C.M. Kao, M.J. Wu, J. Hazard. Mater. 74 (3) (2000) 197.

[117] C.K.J. Yeh, Y.A. Kao, C.P. Cheng, Chemosphere 46 (1) (2002) 67.

[118] C. Liang, C.J. Bruell, M.C. Marley, K.L. Sperry, Chemosphere 55 (2004)

1213.

[119] USEPA, NATO/CCMS Pilot Study, Evaluation of Demonstrated and

Emerging Technologies for the Treatment of Contaminated Land and

Groundwater (Phase III), 1998 Special Session, Treatment Walls

and Permeable Reactive Barriers, Number 229, EPA 542-R-98-003,

North Atlantic Treaty Organization, Vienna, Austria, 1998.

[120] H.-N. Lim, H. Choi, T.-M. Hwang, J.-W. Kwang, Water Res. 36 (2002)

219.

[121] Y. Zeng, P.K.A. Hong, D.A. Wavrek, Environ. Sci. Technol. 34 (2000)

854.

[122] A. Goi, N. Kulik, M. Trapido, Chemosphere 63 (10) (2006) 1754.

[123] C.D. Johnston, J.L. Rayner, B.M. Patterson, G.B. Davis, J. Contam.

Hydrol. 33 (1998) 377.

[124] T.F. Guerin, S. Horner, T. McGovern, B. Davey, Water Res. 36 (2002) 15.

[125] A.D. Cohen, M.S. Rollings, W.M. Zunic, J.D. Durig, Water Resour. Res.

25 (9) (1991) 1047.

[126] R. Naidu, R.S. Kookana, D.P. Oliver, S. Rogers, M.J. Mclaughlin,

Contaminants and the soil environment in the Australian-Pacific

Regions. Proceedings of the Australasia-Pacific conference on contami-

nants and soil environment in the Australasia-Pacific regions, held in

Adelaide, Kluwer Academic Publishers, Australia, 18–23, 1996.

[127] J.M. Duran, J. Vogan, J.R. Stening, Reactive Barrier Performance in a

Complex Contaminated and Geochemical Environment. II International

Conference on Remediation of Recalcitrant Compounds, Montery, CA,

Columbus, May 22–25, 2000.

[128] B.M. Patterson, M.E. Grassi, B.S. Robertson, G.B. Davis, A.J. Smith,

A.J. Mckinley, Environ. Sci. Technol. 38 (2004) 6846.

[129] L.A. Schipper, M. Vojvodic-Vukovic, Water Res. 35 (14) (2001)

3473.

[130] B.M. Patterson, P.D. Franzmann, G.B. Davis, J. Elbers, L.R. Zappia, J.

Contam. Hydrol. 54 (2002) 195.


Top Related