You are on page 1of 13

ARTICLE IN PRESS

WAT E R R E S E A R C H 42 (2008) 2605 – 2617

Available at www.sciencedirect.com

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

Treatment of brackish produced water using carbon


aerogel-based capacitive deionization technology

Pei Xu, Jörg E. Drewes, Dean Heil, Gary Wang


Environmental Science & Engineering Division, Colorado School of Mines, Advanced Water Technology Center (AQWATEC),
Golden, CO 80401-1887, USA

art i cle info ab st rac t

Article history: Capacitive deionization (CDI) with carbon-aerogel electrodes represents a novel process in
Received 27 September 2007 desalination of brackish water and has merit due to its low fouling/scaling potential,
Received in revised form ambient operational conditions, electrostatic regeneration, and low voltage requirements.
8 January 2008 The objective of this study was to investigate the viability of CDI in treating brackish
Accepted 10 January 2008 produced water and recovering iodide from the water. Laboratory- and pilot-scale
Available online 20 January 2008 experiments were conducted to identify ion selectivity, key operational parameters,
evaluate desalination performance, and assess the challenges for its practical applications.
Keywords:
The performance of the CDI technology (CDTs) system tested was consistent throughout
Carbon aerogel
the laboratory- and field-scale experiments. Deterioration of the carbon-aerogel electrodes
Capacitive deionization
was not observed during testing. The degree of ions adsorbed to the carbon aerogel (in mol/g
Desalination
aerogel) during treatment of brackish water was dependent upon initial ion concentrations
Brackish water
in the feed water with the following selectivity I4Br4Ca4alkalinity4Mg4Na4Cl. The
Produced water
preferential sorption of iodide revealed merit to efficiently recover iodide from brackish
Water reuse
water even in the presence of dominant co-ions. The research findings derived from this
study identified parameters that merit further improvements regarding design and
operation, including modification of pore-size distribution of aerogel, development of high
capacitance and low-cost electrode materials, reducing the dead volume after regeneration
and rinsing, minimizing energy consumption, and maximizing system recovery.
& 2008 Elsevier Ltd. All rights reserved.

1. Introduction (IOGCC/ALL Consulting, 2006; ALL Consulting, 2003; Veil et al.,


2004). Treating produced water for beneficial use is being
Due to population growth, reductions in fresh or traditional investigated as an attractive alternative to traditional disposal
water supplies, growing vulnerability due to droughts, methods, such as surface discharge or deep well injection
competition for conventional water sources, development of (IOGCC/ALL Consulting, 2006; Veil et al., 2004) offering an
unconventional water supplies (such as wastewater, brackish opportunity to augment water supplies in areas that lack
water, and seawater) are increasingly promoted worldwide. In fresh water supplies. Produced water is characterized by total
the US, a large amount of brackish water, termed produced dissolved solids (TDS) concentrations exceeding 3000 mg/L
water, is generated during natural gas operations (Benko and and desalination is often required prior to surface discharge
Drewes, 2008). Produced water management is critical to a and beneficial use of produced water. In addition, some
sustainable natural gas operation because of environmental produced waters are characterized by an elevated concentra-
concerns, high disposal costs, and limited disposal options tion of valuable elements, such as iodide. Iodine is an

Corresponding author. Tel.: +1 303 273 3401; fax: +1 303 273 3413.
E-mail addresses: pxu@mines.edu (P. Xu), jdrewes@mines.edu (J.E. Drewes), dheil@mines.edu (D. Heil), ljwang@mines.edu (G. Wang).
0043-1354/$ - see front matter & 2008 Elsevier Ltd. All rights reserved.
doi:10.1016/j.watres.2008.01.011
ARTICLE IN PRESS
2606 WAT E R R E S E A R C H 42 (2008) 2605– 2617

essential and rare element with increasing demand in many solved impurities and the electrodes also contribute to the
industrial applications. Recovery of iodide from produced separation mechanisms of the electrodes. For instance,
water provides an additional economic benefit to produced functional groups such as carbonyl and phenolic groups on
water management by off-setting treatment costs and mini- the carbon electrode can react with cations and form
mizing brine volume for disposal. chemical bonds (Yang et al., 2001). A decrease in the specific
Today, ion-exchange (IX) resins and membrane technolo- adsorption between the electrodes and the ions is a promis-
gies are the most commonly used methods to deionize ing way to eliminate the affinity of the ions held on the
produced water. Produced waters with a predominant sodium electrodes during the regeneration step, increase the regen-
biocarbonate make-up are frequently treated by the Higgins eration efficiency, and enhance CDI performance. Carbon
Loop continuous IX process (IOGCC/ALL Consulting, 2006). aerogel also exhibited high electrosorption capacity to certain
The Higgins Loop IX process is achieving high water ions such as iodide (Ying et al., 2002), thus can be used
recoveries and is cost-efficient only in early field operations to selectively recover valuable constituents from water
with high water volumes (as water volumes decrease in a field solutions.
development and the water to the plant diminishes, the fixed Previous research demonstrated that the efficiency of CDI
costs drive up the cost per volume) (Matthews, 2007). A strongly depends upon the surface property of electrodes
significant cost disadvantage of IX processes is the (a) use of such as surface area, microstructure and size distribution of
hydrochloric acid or sulfuric acid for resin regeneration that pores, chemical functional groups, and adsorption properties
needs to be delivered to remote sites, (b) the generation of a (Pekala et al., 1998; Gabelich et al., 2002; Yang et al., 2001;
sodium-heavy brine that must be hauled over difficult terrain/ Hou et al., 2006). Carbon aerogel is the most common
dirt roads, and (c) the need for wastewater injection disposal electrode material in CDI treatment. It is a special class of
into class III disposal wells. Since the Higgins Loop IX process organic aerogels with unique porous structure consisting of
is limited to treatment of sodium bicarbonate type of water interconnected nanometer-sized particles (3–30 nm) and
only, and produced water streams are dominated by sodium small interstitial pores (o100 nm) (Pekala et al., 1998; Tran
chloride, these waters are not cost-efficiently treatable by this et al., 2003). The monolithic structure, high electrical con-
process. Membrane technologies including reverse osmosis ductivity (10–100 Siemens/cm), high specific surface area
and electrodialysis have been applied to treat produced (400–1000 m2/g), and controllable pore size distribution of
waters with different water compositions, but are typically carbon aerogels make them ideal as electrode materials for
still in pilot testing (IOGCC/ALL Consulting, 2006; Funston electrochemical double layer capacitors and electrosorptive
et al., 2002; Spitz, 2003; Xu et al., 2008). While membrane processes including CDI (Farmer et al., 1996; Pekala et al.,
treatment is also a highly efficient desalination technology, it 1998; Ying et al., 2002). Besides carbon aerogel, other
requires pre-treatment, such as conventional coagulation/ materials are also being investigated as CDI electrodes, such
flocculation, followed by granular media filtration or micro- as nanoporous activated carbon cloth (Oh et al., 2006), titania-
filtration/ultrafiltration. The complexity of these treatment incorporated activated carbon cloth (Ryoo et al., 2003), carbon
processes, large volumes of waste generated during opera- aerogel–silica gel composite electrodes (Yang et al., 2005),
tion, sensitivity of membranes, and low tolerance to system graphitized-carbon monolithic column (Hou et al., 2006), and
upsets (due to well water quality changes, temperature multiwall carbon nanotubes (Zhang et al., 2006; Bordjiba
variations, filter plugging) have hindered a widespread et al., 2007).
application of membrane treatment in the field of produced The CDI process exhibits several advantages such as a
water treatment. New desalination technologies need to be simple, modular design, no need for a high-pressure pump or
evaluated to overcome the limitations of existing desalination heater, as well as operation at ambient conditions and low
processes, promoting beneficial use of produced water. voltages. The CDI process removes ions by charge separation
Capacitive deionization (CDI) with electrodes represents a and the polarity of the electrodes can be reversed between
novel process in desalination of brackish water or production separation cycles. The scaling problems commonly asso-
of high-grade deionized water. CDI can be viewed as an ciated with membrane processes therefore may be avoided.
electrochemical process that takes place in a ‘‘flow through’’ Unlike ion exchange, no additional chemicals are required for
double layer capacitor. During CDI treatment, water is flowing regeneration of the electrosorbent in this system. The down
between the porous surface electrode pairs usually having a time of the system due to cleaning and maintenance is
potential difference of 1.0–1.6 direct current (DC) voltage. The expected to be low. Because of the low fouling and scaling
controlling separation mechanism of CDI is the electrostatic potential, the pretreatment for CDI is minimum (e.g.,
attractive force between the ionic species in the solute and cartridge filtration to remove particles), and the chemicals
the charged electrodes. The negative electrodes attract required for scaling control and cleaning are optional and site
positively charged ions such as calcium, magnesium, and dependent. Given the fact that CDI is only effective in the
sodium. Conversely, the positive electrodes attract negatively removal of charged constituents from water, post-treatment
charged ions such as chloride, nitrate, and sulfate, producing is often required to remove non-ionic organics and pathogens
deionized water. After the electrodes are saturated with salts to meet product water quality standards. As an emerging
or impurities, the electrodes are regenerated by eliminating desalination technology, CDI is still in the developing stage.
the electric field. The adsorbed ions are desorbed from the Most of the studies previously reported on CDI are theoretical
surface of the electrodes, and released into a relatively small, and experimental in nature. Research is needed to gain a
concentrated regenerant stream. In addition to the electro- better understanding of CDI’s potential as a practical desali-
static attractive force, Faradaic reactions between the dis- nation technology.
ARTICLE IN PRESS
WAT E R R E S E A R C H 42 (2008) 2605 – 2617 2607

This study investigated the viability and ion selectivity of

0.02470.006
CDI in treating brackish water generated during natural gas

2.170.5
Alkalinity 235720 mg/L as CaCO3

Sr
operation for beneficial water use and iodide recovery.
Laboratory- and pilot-scale tests were conducted to identify
key operational parameters, optimize system design and
operation, evaluate treatment performance, and critically

4.070.45 L/m mg
Specific UVA
0.09670.021
assess benefits and technical challenges of CDI in practical

0.3970.06
2.770.6

5078
applications.

Si

I
2. Materials and methods

97.8714.2
22507327
2.1. Water quality

Na
Water tested during this study was collected from a natural
gas operation site in Eastern Montana, USA. A comprehensive

0.001370.0005
water analysis was conducted to identify water quality

Total hardness 124723 mg/L

0.0770.03
constituents in produced water critical to CDI treatment and

Mn
beneficial use. Analytical methods and sampling strategies
followed protocols described in standard methods (APHA,

as CaCO3
AWWA, WEF, 2005) or published elsewhere (Xu and Drewes,
2006). Table 1 presents average concentrations of major

0.7070.41 mg/L
Oil and grease
0.45770.078
constituents (in mg/L and mmol/L) from 22 grab samples

0.6470.21
11.171.9

51717
collected over 4 years from the gas production well.

Mg

Br
The produced water was characterized as a brackish
groundwater of sodium chloride type with a pH of 8.457
0.22. The TDS concentration was quantified as 55207718 mg/L
Table 1 – Average water quality of produced water extracted from a natural gas operation site

with a specific conductance of 10,5107934 mS/cm. Besides

0.17670.028
6.971.1
sodium and chloride, major constituents with concentrations TDS 55207718 mg/L

K
of less than 100 mg/L were calcium, magnesium, bromide,
and iodide. The well water was elevated in iodide with a
concentration of 5078 mg/L. The water was classified as hard
(total hardness of 124723 mg/L as CaCO3) with an alkalinity of

10.074.3 m1
0.73870.133
29.575.3

UVA-254
235720 mg/L as CaCO3. Minor constituents with concentra-
Ca

tions of less than 10 mg/L were aluminum, boron, barium,


potassium, silicon, and strontium. No other inorganic con-
stituents were detected in the well water above the detection
limits of the analytical methods applied.
0.014670.004

93.3724.1
33067854

The total organic carbon (TOC) concentration in the


2.070.5
Ba

Cl

well water was 1.7570.20 mg/L, and organics were character-


10,5517934 mS/cm

ized by moderate to high aromaticity (specific UV absor-


Conductivity

bance (i.e., calculated as the ratio between UVA-254 and


DOC) equaled 4.070.45 L/mg m). Oil and grease were
1.7570.20 mg/L
0.35270.028

detected at low concentrations. Hydrocarbons and BTEX


3.870.3

compounds (i.e., benzene, toluene, ethylbenzene, m, p-xylene,


TOC
B

o-xylene), however, were not detected in the well water


samples.

2.2. Carbon aerogel and aquacell assembly


pH 8.4570.22

0.004170.0078
0.1170.21

The carbon aerogel was manufactured by the CDT Systems,


Al

Inc. (Dallas, TX), using a resorcinol/formaldehyde polymer-


ization and pyrolization process. Carbon veil was soaked with
a formaldehyde/resorcinol resin in a polypropylene mold. The
polymerization process was completed in a temperature-
Analytes conc.

controlled oven at 85 1C for 48 h. The sheets were removed


Conc. (mg/L)

Conc. (mg/L)

from the molds and solvent washed to eliminate any


Organics
Cations

Anions

impurities and retained water. The air-dried polymerized


Conc.
mM

mM

sheets were then stacked in a furnace at 1000 1C for 72 h in a


continuous nitrogen flow environment. After pyrolization, the
ARTICLE IN PRESS
2608 WAT E R R E S E A R C H 42 (2008) 2605– 2617

carbon aerogel sheets were cut to desired size, stacked, and 2.3. Laboratory bench-scale testing
included as electrodes in a specially machined polypropylene
housing. Each carbon aerogel pair was separated by con- Bench-scale CDI testing units used in this study consisted of
ductive carbon spacers and a non-conductive glass fiber two carbon aerogel-based cells provided by CDT Systems, Inc.
screen to prevent short-circuiting. The distance between each The CDTs testing unit held 1.3 L of water and had a dry
pair of carbon-aerogel sheets was 2.3 mm (about 90 mil). weight of 4 kg. Each cell contained 24 sheets of carbon aerogel
A stack of carbon-aerogel sheets, spacers, and their asso- (12 cathodes and 12 anodes) with dimensions of 159 
ciated electrical bus connections were pressurized in a 286  0.81 mm3 each, weighing about 650 g of total aerogel.
polypropylene housing. No epoxy was used to assemble the Synthetic brackish water or produced water was fed into
aquacell. Four pairs of graphite terminals provided connec- the cells by a Cole-Parmer Masterflex L/S pump (Barrington,
tion to a power supply and measurement points to check the IL). Prior to testing, the produced water fed to the units was
voltage during charging and discharging. processed through a 5 mm cartridge and a 0.45 mm filter bag
The carbon aerogel had average density of 0.78 g/cm3, (Cole-Parmer, Vernon Hills, IL). The required power to charge
bulk resistivity of 20 m ohm cm, and specific capacitance the cell was provided by a BK Precision 1746A DC power
of 2 Farad/cm2 (Welgemoed and Schutte, 2005). The Bru- supply (Yorba Linda, CA). Bench-scale CDTs cells were tested
nauer–Emmet–Teller (BET) surface area and pore size dis- with a focus on system stability, regeneration effectiveness,
tribution were measured using a five-point N2 gas adsorption ion selectivity, and desalting efficiency.
technique (ASAP 2020; Micromeritics, Norcross, GA). The Experiments were initiated with the assessment of ion
average pore size and pore size distribution were determined removal properties of the cells with synthetic solutions, then
from desorption of N2 according to a theory developed by progressed to the treatment of a synthetic surrogate of the
Barrett, Joyner, and Halenda (BJH) (Barrett et al., 1951). produced water, and culminated in the treatment of produced
Because the carbon-aerogel material was very hydrophilic water from the natural gas production site.
(with no measurable contact angle using sessile drop Prior to each experiment, the cells were first equilibrated
method), the samples were pretreated at 250 1C in a vacuum with the tested feed water by flushing the cell until the
environment for 3 h to remove adsorbed water and carbon effluent electrical conductivity (EC) was equal to the EC of the
dioxide gas prior to measurement. The carbon-aerogel sheet influent. Once equilibration was achieved, a voltage of 1.3 V
samples were characterized to have an average BET surface was applied to the cell with the pump off to pre-charge the
area of 113 m2/g, a BJH desorption pore volume of 0.177 cm3/g, cell for 10 min. The treatment phase was initiated and the
and a BJH desorption average pore size of 4.28 nm. The effluent was collected in 5 min intervals for EC and other
incremental and cumulative pore area of the carbon aerogel measurements. Except as indicated specifically, regeneration
as a function of the pore size is shown in Fig. 1. About 96.2% of was accomplished by applying reverse polarity for 15–18 min
the pores in carbon aerogel had sizes between 2.65 and to discharge the cell with 2 L of fresh feed water flushed
7.69 nm, and 41.8% of the pores had sizes between 3.0 through the cell for 1 h. After the regeneration step of each
and 3.4 nm. cycle, the cell was rinsed for 10 min with feed water at the

180 100
Cumulative
160 90
Incremental

140 80
Cumulative pore area (m2/g)

Incremental pore area (m2/g)

70
120
60
100
50
80
40
60
30
40
20

20 10

0 0
20 40 60 80 100 120 140
Pore size (A)

Fig. 1 – Pore size distribution of carbon aerogel.


ARTICLE IN PRESS
WAT E R R E S E A R C H 42 (2008) 2605 – 2617 2609

same flow rate, to bring the EC in the cell to near that of the performance using the CDTs pilot-scale testing unit in a
feed water. The cell was then pre-charged to initiate the next laboratory setting.
treatment cycle.

3. Results and discussion


2.4. Pilot-scale field testing
3.1. Operational stability of CDTs cells
In order to verify findings derived from bench-scale experi-
ments, a field evaluation with two pilot-scale CDTs testing The reproducibility and operational stability of the CDTs cells
cells was conducted at a gas operation site in Montana. The were assessed by treating a synthetic feed water containing
consistency of bench- and pilot-scale testing regarding design 5000 mg/L sodium chloride and 50 mg/L iodide added as
and operation would allow some assessments regarding up- potassium iodide. The bench-scale cell was operated
scaling of the CDTs process to full-scale treatment plants. in a single pass continuous-flow mode with a flow rate of
The pilot-scale CDTs test units were approximately 10 250 mL/min for 50 min treatment. A total of five cycles were
times larger than the bench-scale units, containing 6.2 kg of completed in succession. Mass balance of sodium chloride
aerogel each. Water was initially delivered from the wellhead was based on measurements of EC at 5 min intervals in
into a feed water reservoir consisting of three open drums of samples taken during the treatment phase, during the final
2000 L. The three drums allowed dissipation of methane gas solution for the regeneration phase, and in the combined
prior to contact with the CDTs cell and served as a short-term solution from the rinse phase of each cycle. The concentra-
storage reservoir. The water was then pumped out of the tion of both iodide and iodine were measured in 15 min
reservoir through a 5.0 mm filter followed by the CDTs units intervals for the effluent from the treatment phase, in the
(Fig. 2). After preliminary testing, the optimum treatment final solution from the regeneration phase, and the combined
cycle was determined and consisted of 80 min of treatment solution from the rinse phase.
(ion removal) followed by 20 min of regeneration, and 10 min The effluent conductivity and iodide concentration over five
of purging using filtered raw water prior to the next treatment consecutive sorption/desorption cycles are presented in Fig. 3.
stage. The cells were charged separately by two DC power The asymmetrical nature of the treatment and regeneration
sources at a voltage of 1.3 V and a current varying from 60 to curves is characteristic of the different kinetics associated
70 A. The exact voltage and amperage were recorded to with ion sorption onto and desorption from the carbon-
determine power consumption. Electrical conductivity was aerogel surface. Iodide exhibited a different adsorption
measured in the inlet and outlet from each cell using in-line kinetic as compared with chloride (in terms of conductivity).
electrodes, and EC data were logged by a laptop computer Iodide concentrations were still at a minimum level, while
using a customized Labview program (National Instruments, sodium chloride species were reaching feed water concentra-
Inc.). In addition, influent and effluent samples were collected tion levels in the cell effluent. The effluent conductivity
at 6 min intervals and analyzed for iodide using an ion- reached a minimum of 8.5 mS/cm after 15 min of each
selective electrode and pH, conductivity, and temperature. treatment phase, after which the effluent conductivity slowly
The samples were also shipped to the Colorado School of approached feed water levels (10.2 mS/cm). Iodide concentra-
Mines laboratory for further analysis including TOC, hard- tion decreased from an initial concentration of 50 mg/L in the
ness, alkalinity, UV absorbance, cations, and anions. influent to 17–25 mg/L in the effluent during the sorption
After the field testing in Montana, additional experiments phase. The iodide retention was above 50%, as compared with
were conducted using synthetic water to investigate the only 5–6% removal of sodium chloride. The average concen-
effect of flow rate and feed water concentration on sorption tration of dissolved iodine generated in the effluent varied

Data collection and analysis

Aquacell #101 Aquacell #102

Feed
Pump water
reservoir

DC power supplies DC power supplies 5µm cartridge


filter

Fig. 2 – Flow schematic of CDTs pilot-scale test system employed during field trials.
ARTICLE IN PRESS
2610 WAT E R R E S E A R C H 42 (2008) 2605– 2617

14 140 higher relative error than the laboratory-scale experiments,


One cycle Conductivity
12 120 likely due to a fluctuation of feed water quality and challenges
of operating larger-scale testing units at a remote site.
10 100
EC (mS/cm)

Iodide (mg/L)
8 80 3.2. Maximum sorption capacity of bench-scale and pilot-
6 60 scale cells
4 Iodide 40
To measure the maximum sorption capacity of the CDTs
2 20 cells, batch equilibrium experiments were conducted in a
0 0 continuous-recycling regime. For bench-scale cells, 2 L of
0 100 200 300 400 500 600 700 synthetic sodium chloride solution (2000 mg/L) was pumped
Time (min) through the cells at a flow rate of 250 mL/min and an applied
Fig. 3 – Reduction in EC and iodide over five consecutive power of 1.3 V. The conductivity of the solution in the feed
cycles (laboratory-scale experiments: single pass reservoir was monitored continuously until steady-state
continuous-flow operation at 250 mL/min with feed water conductivity was achieved. The total capacity of the cells
composition of 5000 mg/L NaCl and 50 mg/L iodide). was determined by mass balance calculations. The maximum
sorption capacity of the cells was determined to be 7.0 mg
NaCl/g carbon aerogel. For the pilot-scale cells, 38.75 L of
produced water (5200 mg/L TDS, other water quality para-
between 1.1 and 5.7 mg/L. Given the half-cell potential for meter listed in Table 1) was pumped through the cells at a
I2/2I of 0.536 V, a portion of iodide ions was reduced to flow rate of 250 mL/min and an applied power of 1.3 V. The
iodine at the applied potential of 1.3 V during CDI treatment: maximum sorption capacity of the cells was 7.1 mg TDS/g
2HI-H2+I2. The depletion of H+ ions was consistent with the carbon aerogel. Thus, the bench- and pilot-scale CDTs cells
increase of effluent pH from 8.3 to 9.4. exhibited almost the same sorption capacities.
The final concentration of iodide in the regenerant water
varied between 71 and 89 mg/L, while the final conductivity in 3.3. Effect of flow rate and initial feed concentration on
the regenerant water was between 11.9 and 12.2 mS/cm. For electrosorption performance
each cycle, 77–107% of adsorbed iodide was recovered from
the carbon aerogel during the regeneration phase. (Note: a The effects of flow rate and initial feed concentration on
recovery greater than 100% was due to the desorption of electrosorption performance were investigated using syn-
iodide from the previous run.) During the same time, the thetic sodium chloride solutions at a single-pass continuous
regeneration cycle could only remove between 68% and 94% flow regime (Fig. 4). At the feed flow rate of 250 mL/min using
of adsorbed sodium chloride from the carbon aerogel. the bench-scale testing unit, the increase of dissolved salts in
To illustrate the reproducibility between each treatment feed stream resulted in an initially linear increase of
and regeneration cycle, the relative standard deviation (RSD) operational sorption capacity (Fig. 4(a)). After the feed
regarding minimum effluent conductivity was 1.3%. The RSD concentration was greater than 2000 mg/L, the operational
regarding minimum effluent iodide was 14.7%. The higher sorption capacity of the electrodes leveled off reaching 6 mg
relative testing deviation of measured iodide data was likely TDS per gram of carbon aerogel. The average sorption rate
due to the longer sampling intervals (15 min versus 5 min for followed the same trend: increasing initially and stabilized at
EC). Error associated with measuring solution conductivity 0.1 mg TDS per gram of carbon aerogel per minute of
and iodide is expected to be in the same range. treatment time (Fig. 4(a)).
The potential for short-term decline in sorption capacity of The effect of flow rate on electrosorption of CDI was
carbon aerogel was studied over five cycles. The minimum examined using the pilot-scale testing unit treating synthetic
observed sodium chloride concentration in the effluent sodium chloride solutions. As compared with the feed
increased by 2.8%. The total sodium chloride adsorbed to concentration, the increase in feed flow rate exhibited a
carbon aerogel decreased gradually by 25.6% between the first limited effect on the electrosorption performance of the
and fifth cycle. However, the accumulated sodium chloride in electrodes (Fig. 4(b)). At the feed concentration of 500 mg/L,
the cell over the five cycles was recovered nearly completely the operational sorption capacity of the electrodes decreased
after an overnight (16 h) regeneration cycle. The decline in from 3.2 to 2.5 mg TDS per gram of carbon aerogel when the
sorption capacity during consecutive runs is likely due to slow flow rate increased from 700 to 3000 mL/min. The average
kinetics of ion desorption from the carbon surface. No sorption rate, however, increased from 0.013 to 0.052 mg TDS
appreciable decline in carbon-aerogel performance was ob- per gram carbon aerogel per minute with the increase of flow
served for iodide over the five consecutive cycles, which rate. This implies that high flow rate should be employed for
might be due to the fact that electrosorption equilibrium industrial CDI process design and operation since fewer CDI
of iodide/iodine was not achieved over the five successive modules will be required to produce the same amount and
cycles. quality of desalinated water than at low flow rate.
Throughout the laboratory- and pilot-scale testing, each To better understand the effect of feed flow rate and
experiment was conducted by applying 2–5 consecutive concentration on electrosorption performance of the electro-
cycles under the same operational conditions. The pilot-scale des, the experimental results were plotted as a function of
experiments treating produced water in the field exhibited TDS loading—mass of TDS loaded per gram of aerogel per
ARTICLE IN PRESS
WAT E R R E S E A R C H 42 (2008) 2605 – 2617 2611

7.0 0.20 3.5 0.10


Sorption (mg TDS/g aerogel)

Sorption (mg TDS/g aerogel)


6.0 3.0

Sorption rate (mg TDS/g


Sorption rate (mg TDS/g
0.08
5.0 0.15 2.5

aerogel/min)
aerogel/min)
2.0 0.06
4.0
0.10
3.0 1.5 0.04
2.0 0.05 1.0
0.02
1.0 Sorption Sorption rate 0.5 Sorption Sorption rate

0.0 0.00 0.0 0.00


0 2000 4000 6000 0 1000 2000 3000 4000
Feed TDS concentration (mg/L) Flow rate (mL/min)

Fig. 4 – Effect of flow rate and feed concentration on operational sorption capacity and average sorption rate of the carbon
aerogels using synthetic water. The error bars represent the standard deviation of 3–5 runs of each testing condition. (a) Feed
flow rate 250 mL/min using laboratory bench-scale unit; (b) feed TDS 500 mg/L using pilot-scale testing unit.
Sorption rate (mg TDS/g aerogel/min)

0.14 micropores, resulting in lower electrosorption capacity (Yang


et al., 2001; Gabelich et al., 2002). The electrical double layer
0.12
thickness is primarily affected by ion solution concentration
0.10 and applied voltage. Farmer et al. (1996) reported that the
electrical double layer thicknesses were 1 and 20 nm for 0.1
0.08
and 104 M electrolytes solutions, respectively. Our study
0.06 confirmed that increasing electrolytes loading to the elec-
0.04 trode surfaces through increasing feed concentration could
Laboratory testing with NaCl solution
reduce electrical double layer thickness, and consequently,
0.02 Field testing with produced water
improve electrical capacity of the carbon aerogel. However,
0.00 the mass transfer limitations may still be in effect due to
0 0.5 1 1.5 2 micropores overlapping and monolayer sorption behavior of
TDS loading (mg TDS/g aerogel/min) carbon-aerogel electrodes. These effects may explain the
observed trend that the sorption rate increased with initially
Fig. 5 – Effect of TDS loading on sorption rate of carbon
increasing electrolytes loading but leveled off after the
aerogel. The error bars represent the standard deviation of
loading rate reached 0.55 g TDS/(g aerogel min) as illustrated
3–5 runs of each testing condition.
in Fig. 5.
No significant correlation between iodide sorption and flow
rate was observed during produced water treatment. Within
minute during the sorption stage (i.e., feed flow rate  feed the tested flow rate ranging from 600 to 1500 mL/min, the
concentration/mass of the electrodes) (Fig. 5). At lower feed retained iodide in the carbon-aerogel electrodes varied
concentrations (i.e., lower TDS loading), the sorption between 0.13 and 0.15 mg I/g aerogel and the sorption rate
rate increased linearly. When TDS loading exceeded varied in the range of 0.004–0.006 mg I/(g aerogel min).
0.55 g/(g aerogel min), the sorption rate leveled off at approxi-
mately 0.09 mg TDS/(g aerogel min). 3.4. Ion selectivity and effect of temperature on
Previous studies have demonstrated that the adsorption electrosorption performance
capacity of the carbon aerogels is mainly affected by electrical
double-layer capacity due to the electrostatic attractive force Ion selectivity of the carbon aerogel was examined using pre-
between the ions and the electrode (Yang et al., 2001; Gabelich filtered produced water at ambient room temperature of 23 1C
et al., 2002; Ying et al., 2002). In porous electrodes such as and temperature of 12 1C, considering in-situ well water
carbon aerogel, the electrical double layers are formed inside conditions. Laboratory-scale batch treatment was conducted
the pores instead of adjacent to the electrode surface. The at each temperature and four treatment cycles were com-
pores can greatly increase the effective surface area of the pleted. During the treatment phase, 6.25 L of produced water
porous electrodes as well as the electrical capacity. However, was re-circulated at a flow rate of 250 mL/min for a period of
when the pore size is of similar magnitude as the electrical 50 min. For the experiments conducted at 12.5 1C, the
double layer thickness, the electrical double layer inside the temperature of the water discharged from the CDTs bench-
pores overlaps, resulting in a loss of electrical capacity (Yang scale cell was checked periodically and ranged between 11.6
et al., 2001). This overlapping effect exists in microporous and 13.2 1C.
(o2 nm) and part of mesoporous (2–50 nm) regions. Given the During CDI treatment, reduction in EC at 23 1C was very
microporous structure of the carbon aerogel (see Fig. 1), this similar to that achieved at 12.5 1C and resulted in reduction
overlapping effect likely prevents ions from entering the between 6.7% and 7.0% (Fig. 6). Iodide reductions in both the
ARTICLE IN PRESS
2612 WAT E R R E S E A R C H 42 (2008) 2605– 2617

treated water and regenerant water were also similar at the water is consistent with results reported by Gabelich et al.
two temperatures examined and varied between 66% and (2002) in treating Colorado River Water.
69%, with iodide concentrations close to 15 mg/L in the Previous studies suggested that, for ion species having
treated water and close to 90–100 mg/L in the regenerant similar initial solution concentration (in terms of molarity),
solution. Divalent cations, such as Ca, Mg, and Sr, exhibited the hydrated radius might control their sorption capacity of
between 21.7% and 22.5% removal on average at 23 1C and carbon-aerogel electrodes (Gabelich et al., 2002; Ying et al.,
between 16.3% and 18.0% at 12.5 1C, respectively. The removal 2002). Monovalent ions such as sodium with smaller hydrated
of sodium, however, was higher at 12.5 1C than at room radii were preferentially removed from solution over multi-
temperature (13.3% versus 8.2% on average). The removal of valent ions (such as calcium) on a percent or molar basis.
potassium was similar to the divalent cations, 21.3% and There was also evidence that the counter ion present could
15.7% at 23 and 12 1C, respectively. The carbon aerogel did not play an important role in an individual ion’s sorption capacity
remove any silica and boron from the produced water likely (Gabelich et al., 2002; Ying et al., 2002). In this study, it was
due to the undissociated state of these constituents within observed that, in a competitive multi-ionic solution, feed
the operating pH range (8.3–8.5). Increasing the pH to above concentration seemed to play a more important role in ion
9.5, boron will dissociate and could be removed by the CDTs uptake than the ionic hydrated radius. For instance, potas-
process potentially meeting water quality standards for sium has an effective ionic radius of 3.31 Å in aqueous
beneficial use such as irrigation. solution, much smaller than sodium (3.58 Å), calcium
The amount of adsorbed ions by carbon aerogel (in mol/g (4.12 Å), and magnesium (4.28 Å). Therefore, it is expected
aerogel) in treating produced water increased in the order of that the sorption capacity of potassium should be higher than
Na (7.9  105)bCa (1.6  106)4Mg (9.7  107)4K (4.6  107) sodium, calcium, and magnesium based on hydrated radius.
for cations, and Cl (7.7  105)bBr (3.6  106)4I (2.6  106) The experimental results, however, showed that the mass of
for anions (Table 2). This sorption trend in treating produced adsorbed ions was dependent upon the feedwater concentra-
tion (in terms of molarity) rather than the hydrated radius,
and followed the order of NabCa4Mg4K. A similar conclu-
sion can be drawn for anions. Ying et al. (2002) reported that
80 the sorption capacity of anions followed the trend of Br4Cl
considering similar feed concentrations (in terms of molar-
70
23 degree C ity). This study observed the opposite trend of sorption
60 12.5 degree C capacity, which corresponded to the initial feed concentration
Removal (%)

50 (in terms of molarity) of Cl4Br (Table 2).


The removal of iodide in the produced water reached 69%,
40
much higher than other ions (Table 2). Ying et al. (2002)
30 attributed high sorption capacity of iodide to its higher partial
20 charge-transfer coefficient than chloride and bromide. The
partial charge-transfer coefficient is an indicator of how many
10
electrons can be released from the adsorbate to an adsorbent.
0 In this study, a concentration between 1.1 and 5.7 mg/L of
EC I Ca Mg Sr Na K B
iodine was detected in the CDI effluent during produced water
Fig. 6 – Effect of temperature on removal of different treatment. Besides electrostatic adsorption, iodide and iodine
constituents from produced water (batch recycling
operation using laboratory bench-scale unit).

Table 2 – Ion sorption and hydrated area of the carbon aerogel electrodes in treating produced water (batch recycling
operation using laboratory-scale unit at 23 1C)

Ions Hydrated Feed concentration Sorption capacity Removal Hydrated area


radius* (Å) (mM) (  105 mol/g aerogel) (%) (m2/g aerogel)

Na 3.58 97.83 7.9 8.1 19.15


K 3.31 0.18 0.046 21.3 0.10
Ca 4.12 0.74 0.16 22 0.51
Mg 4.28 0.46 0.097 21.7 0.34
Cl 3.32 93.26 7.7 7.6 16.05
Br 3.3 0.63 0.36 50 0.74
I 3.31 0.39 0.26 69.7 0.54

Total cations 8.20 20.09


Total anions 8.32 17.33

Note: *Reference: Nightingale (1959).


ARTICLE IN PRESS
WAT E R R E S E A R C H 42 (2008) 2605 – 2617 2613

can have complex intermolecular interactions with the 3.5. Field performance of CDTs cells
carbon-aerogel material. Iodide and iodine can react with
various carbon groups of the aerogel, such as carbonyl and The sorption and desorption pattern of various constituents
phenolic functional groups, resulting in higher sorption present in the produced water during the 1-stage CDI
capacity. treatment (single pass with two CDTs cells in series) is
To further investigate the effective surface area of the presented in Fig. 7. The sorption and desorption trends of
carbon aerogel in produced water treatment, the amount of each studied constituents during field trials were consistent
aerogel surface area covered by each ion specie (m2/g aerogel) with bench-scale observations. The maximum removal of the
was calculated by multiplying the molar sorption capacity of constituents followed the order of aromatic organic carbon (in
the ions (mol/g aerogel) by Avogadro’s number (6.022  1023) terms of UVabs at 254 nm, 83.3%)4iodide (77%)4Br (62.5%)4
and the two-dimensional area based on the hydrated Ca (40.7%)4alkalinity (as CaCO3, 40.0%)4Mg (34.3%)4Na
radius (m2): (18.4%)4Cl (16.0%). The adsorbed ion mass to carbon aerogel
(in mol/g aerogel) in treating the produced water followed the
Hydrated area ¼ Sorption capacity  Avogadro’s number order of NabCa4Mg4K for cations, and ClbBr4I for anions.
 p  hydrated radius2 . (1) The operational sorption capacities (mol/g aerogel) were
much lower during the field tests than during laboratory
The carbon aerogel exhibited similar sorption capacities of experiments, likely due to low Reynolds numbers during the
total cations and anions, 8.07  105 versus 8.32  105 operation of the pilot-scale CDTs cells. Reynolds numbers
equivalents per gram aerogel (Table 2). Based on the hydrated varied between 55 and 150 for field testing, indicating laminar
radius, cations and anions covered similar surface areas of flow conditions between carbon-aerogel sheets. The laminar
the carbon aerogel, about 20 and 17 m2/g aerogel, respectively. flow may significantly increase the electrical double layer
The utilized surface area was low, about 33% of the total thickness, resulting in low operational electrosorption
surface area as compared with the measured BET surface area capacities.
of 113 m2/g aerogel. Given the average pore size of the aerogel Similar to the observations made during the bench-scale
in the range of 4 nm, the solutes could not effectively diffuse study, TOC concentrations were actually lower in the
into the inner micropores and resulted in lower effective regenerant solution than in the treated effluent (Fig. 7(a)).
surface area for sorption. Gabelich et al. (2002) reported that Adsorption of TOC to the aerogel material during regenera-
only 10% of the carbon aerogel was available for sorption. tion when the cell is uncharged could result in potential

16 120 5000
1.6
Br I Cl
TOC (mg/L), UVA 254 (/cm)

EC UVA254 TOC
14 1.4 100
4000

Cl (mg/L)
I and Br (mg/L)

12 1.2
80
EC (mS/cm)

10 1.0 3000
8 0.8 60
6 2000
0.6 40
4 0.4
20 1000
2 0.2
0 0.0 0 0
0:00 1:12 2:24 3:36 4:48 0:00 1:12 2:24 3:36 4:48
Operating time Operating time

60 3500 10 300
Ca Mg Na
50 3000 8 250
pH and B (mg/L)
Ca and Mg (mg/L)

2500
Alkalinity (mg/L)

40 200
Na (mg/L)

6
2000
30 150
1500 4
20 100
1000
10 2
500 B pH Alkalinity 50
0 0 0 0
0:00 1:12 2:24 3:36 4:48 0:00 1:12 2:24 3:36 4:48
Operating time Operating time

Fig. 7 – Treatment of various constituents in produced water using a 1-stage unit with two cells in series at a flow rate of
560 mL/min and regeneration rate of 1900 mL/min (T ¼ 18.6–27.6 1C).
ARTICLE IN PRESS
2614 WAT E R R E S E A R C H 42 (2008) 2605– 2617

electrode fouling by organic matter by clogging the pores of and even better than the results obtained with the bench-
carbon-aerogel material. UV absorbance exhibited similar scale unit. Using original well water for regeneration, the
sorption and desorption trends as anions during treatment iodide concentration in the regenerant of a third stage could
(Fig. 7(a)). This indicates that aromatic organic acids behave be lower than the influent as a result of dilution.
similarly to anions during the CDI process. The carbon
aerogel exhibited a preferential sorption of iodide from 3.6. CDTs operational performance and sorption efficiency
produced water regarding percent removal. Moreover, the during multi-stage treatment of brackish water (5000 mg/L)
iodide adsorbed onto the electrodes exhibited incomplete
desorption during regeneration in comparison to chloride and Because of the high TDS level in the produced water, the field
bromide (Fig. 7(b)). The initial specific adsorption capacity of experiments employing a 3-stage treatment still could not
iodide to carbon aerogel (without an applied voltage to the meet the water quality standards for beneficial use (Table 3).
electrodes) was determined to be 1.3 mg iodide/g carbon Successive batch experiments were conducted using the
aerogel (10.4 mmol/g carbon). CDTs bench-scale unit to simulate a multi-stage desalination
Calcium, magnesium, and sodium had the same sorption treatment that would reduce the TDS of produced water from
and desorption trend during produced water treatment 5000 mg/L to a level suitable for irrigation (i.e., 1000 mg/L) for
(Fig. 7(c)). The carbon aerogel did not remove any silica and potential beneficial use. Approximately 6.25 L (1.65 gallons) of
retained very little boron from produced water due to the 10 mS/cm synthetic water (5000 mg/L NaCl and 50 mg/L iodide
undissociated state of these constituents within the operating as KI) was fed into the cell at a flow rate of 250 mL/min at
pH range (Fig. 7(d)). The maximum removal of boron was room temperature. Treated product water was collected and
about 16.9% during the first stage of treatment and the boron subsequently used as feed water for the next experiment
concentration in the effluent of the third stage was still simulating a successive multi-stage treatment. Initially,
exceeding 3 mg/L. The pH varied between 7.5 and 9.1 during 10 mS/cm water was used as regenerant water for the first
sorption and regeneration cycles (Fig. 7(d)). The carbon stage and then recycled for the following experiments to
aerogel also exhibited a high retention of alkalinity about reduce regenerant volume and increase system recovery.
40% during produced water treatment (Fig. 7(d)). After regeneration, the cell was rinsed using 10 mS/cm water
At a higher flow rate during regeneration (1700 mL/min for 10 min. The recovery is defined as the percentage of
using prefiltered produced water), the regeneration time was product water based on the total amount of water used for
shortened to one-third of the production time. More than 80% treatment, regeneration, and rinsing.
of adsorbed ions were desorbed from the electrodes during During the 10 cycles of treatment, the conductivity of the
regeneration. However, a rather high volume of brine was water reduced from an initial value of 10 to 5.4 mS/cm. The
produced, for instance, at least the same amount of product incremental reduction in EC (difference between feed water
water for the 1-stage treatment. EC and product water EC) decreased from 1.04 mS/cm for the
To simulate the treatment efficiency of multiple CDTs first cycle to 0.22 mS/cm by the ninth cycle, apparently
stages, the product water (effluent) from the previous stage associated with deficiency of regeneration and rinsing.
was collected and used as influent of the subsequent stage. Improvements in the increment of EC reduction were
The sorption and regeneration cycles by 2-stage and 3-stage observed after overnight regeneration. Iodide concentration
CDTs cells were operated at a flow rate of 550 mL/min. The in the regenerant water was 74.4 mg/L after the first cycle, and
cells were regenerated with filtered well water at a flow rate of then gradually decreased during the first day to 66.4 mg/L in
1700 mL/min. As compared with 1-stage operation, the the recycled regenerant at the end of the fourth cycle. Iodide
treatment efficiencies of salts were constant over the multiple concentration in the regenerant then increased to 89.9 mg/L
stages with 2.8–2.9 mg TDS/g aerogel (Table 3). Iodide recovery after the overnight regeneration, and then decreased to
during the regeneration step was effective, with iodide 50 mg/L at the end of the tenth cycle due to adsorption of
concentrations increasing from 45 mg/L in the influent to iodide onto the carbon aerogel. It is evident that recycling of
values in excess of 120 mg/L in the brine during the regenerant water is detrimental to iodide recovery from
regeneration of the first stage, and in excess of 80 mg/L produced water due to the high affinity of iodide ions to
during the regeneration of the second stage. This was similar carbon aerogel.

Table 3 – Comparison of TDS and iodide removal by 1-, 2-, and 3-stage treatment

Stage Conductivity (mS/cm) Iodide (mg/L)

Infl. Effl. Sorption Infl. Effl. Regenerant Sorption


min. (mg/g aerogel) min. (max.) (mg/g aerogel)

1 11.68 9.15 2.9 45.0 10 123 0.10


2 9.86 7.83 2.8 13.5 3.7 85 0.03
3 6.70 5.33 2.8 6.0 3.5 42–50 0.005
ARTICLE IN PRESS
WAT E R R E S E A R C H 42 (2008) 2605 – 2617 2615

After the tenth cycle, the EC of the feed water was Table 4 – Treatment of 10.0 mS/cm water (5000 mg/L NaCl
decreased in steps to approximately 4.0, 3.0, and 2.0 mS/cm plus 50 mg/L I) using 1.0 mS/cm water to rinse cell
in order to accelerate completion of the experiments. between treatment cycles
This was deemed acceptable because the iodide concentra-
tion in the treated water had stabilized at approximately Cycle EC Increment Increment
10–13 mg/L, and iodide concentration in the regenerant (mS/cm) (mS/cm) (mS/cm)
water was near 50 mg/L. Treatment of the 2.0 mS/cm water (1 mS/cm rinse) (10 mS/cm rinse)
yielded a very low value for the increment of EC reduction of 1 10.0648.70 1.36 1.04
only 0.08 mS/cm. Consequently, with the given experimental 2 7.1046.16 0.94 0.57
set-up at least 10 cycles would be required to decrease EC 3 5.1244.46 0.66 0.42
from 2.0 to 1.0 mS/cm. Because the regenerant water was 4 3.0042.63 0.37 0.15
reused, the EC in the regenerant stage reached values as high
as 13.33 mS/cm. However, after the treated water EC was
below 5.0 mS/cm, EC in the regenerant water tended to
remain near a value of 10.0 mS/cm. This is probably due to
carryover of relatively low EC water remaining in the cell after early stages of the process. Iodide recovery was poor due to
the treatment step. rinsing the cell with low-iodide (I ¼ 10 mg/L)-treated water.
In order to estimate the effect of carryover from regenera-
tion on the treatment of water in the following cycle, the 3.7. Benefits and technical challenges
reduction in operational adsorption capacity was calculated.
It is assumed that the maximum adsorption capacity avail- The performance of the CDTs system was maintained at a
able was 5.54 mg/g aerogel based on the results from the first consistent level throughout the duration of the field trials.
cycle. The residual concentration of sodium chloride in the Scaling or fouling of the aerogel electrodes did not occur
cell after the rinse stage was slightly above 5000 mg/L during treatment of this water source. Although the system
(EC ¼ 10.0 mg/L) throughout the entire range of initial EC as configured did not reduce TDS to the extent needed to
values because produced water was used for rinsing. After meet the targeted water quality standards, the scale at which
draining the cell following the end of the rinse step, at least the test was conducted would allow an extrapolation to
800 mL of this water remained in the cell. Therefore, the determine the number of cells and configuration needed to
adsorption capacity needed to reduce the EC of the residual meet these standards. CDI technology (CDT) was revealed to
water in the cell from 10.0 mS/cm to match the EC of the feed be promising for iodide recovery from the produced water
water at a given stage of treatment was calculated as examined as well as other sources that contain iodide. CDI
needs only simple pre-treatment such as cartridge filtration.
0:80 Lð5000  Ci Þ
Mlost ¼ , The efficiency and production capacity of the system,
650 g
however, need to be improved before CDT will become
where Mlost is the mg of adsorption capacity per g aerogel lost economically feasible to treat water of this level of salinity.
due to carryover, and Ci is the initial concentration of NaCl in The results of the field trials allowed identifying critical
mg/L at a given stage of the treatment process. equipment parameter and operational conditions that merit
Consequently, the percentage of the adsorption capacity further improvement. Development of high capacitance and
available for treating water at a given initial sodium chloride low cost of electrode materials is a key factor for a successful
concentration is and economic commercial application of CDI. By controlling
the pore size distribution of carbon aerogels, the electrosorp-
5:54  Mlost
Mavail ð%Þ ¼  100. tion efficiency of the electrode material can be improved. The
5:54
slow kinetics of transport of ions into and out of the highly
These calculations suggest that the percentage of available porous electrodes should also be addressed. Several efforts
adsorption capacity decreases from 100% at initial EC of have been made to improve the performance of the capacitor.
10.0 mS/cm to only 11% for an initial EC of 2 mS/cm. The Andelman and Walker (2004) used charge-blocking layers
predicted 10-fold decrease in available adsorption capacity is (essentially IX membranes) on the surface of the electrodes to
consistent with the experimentally measured reduction in EC limit migration of ions into and out of the electrodes and
increment from approximately 1 mS/cm at an initial EC of 10 decrease the transition times between purifying and purging
to 0.08 mS/cm at an initial EC of 2 mS/cm. The very low cycles.
increments of EC reduction at low initial EC levels can be The recovery of the CDT process is low for produced water
accounted for based on consideration of the carryover of ions treatment (about 25–33%) due to the substantial amount of
from regeneration. feed water used for electrode regeneration and rinsing. The
To minimize the carryover problem and improve sorption disposal of high volume of brine generated during CDI may be
capacity, 1 mS/cm water was used in the rinse step. As a potential barrier for its practical application. The enhance-
compared to rinsing with 10 mS/cm water, the operational ment of recovery can be achieved by developing faster
sorption capacity of the CDTs cell increased significantly charging and discharging cycles for the electrodes, reducing
using 1 mS/cm water for rinsing (Table 4). The EC of the rinse the volume and time for regeneration and rinsing, as well as
water was increased from 1.0 to 3.5 mS/cm, indicating that reducing dead volume (residual water volume) after regen-
the rinse water could be reused when treating water in the eration.
ARTICLE IN PRESS
2616 WAT E R R E S E A R C H 42 (2008) 2605– 2617

The minimum energy consumption was 0.21 Wh/g salts was derived. The authors also thank Paul Mendell with
removed at applied current of 15 A for both sorption and Mendell Energy Inc., for his financial and technical support,
regeneration. The power consumption of CDI in treating and CDT Systems Inc., for providing CDTs testing units and
produced water could be as high as 0.84 and 0.95 kWh/m3 technical support.
(3.18 and 3.58 kWh/kgal) to meet water quality standards of
1000 and 500 mg TDS/L, respectively. Although applying high R E F E R E N C E S
current could shorten the time for electrode charging and
discharging, the specific power consumption per gram of
removed salt is becoming excessively high. Approximately ALL Consulting, 2003. Handbook on Coal Bed Methane Produced
half of the energy was used for discharging the electrodes Water: Management and Beneficial Use Alternatives. Tulsa,
during the regeneration phase. It implies that the current CDT Oklahoma /http://www.all-llc.com/CBM/pdf/CBMBU/
process is not energy efficient in treating high-salinity water. CBM%20BU%20Screen.pdfS.
Andelman, M.D., Walker, G.S., 2004. US Patent 6,709,560.
Energy consumption can be reduced if energy is retrieved
American Public Health Association (APHA), American Water
during the process of electrode discharge. For example, Shiue Works Association (AWWA), and Water Environment Federa-
et al. (2005) reported a significantly improved efficiency of CDI tion (WEF), 2005. Standards Methods for the Examination of
by using spiral wound electrodes (activated carbon coated on Water and Wastewater, 21st ed.
titanium foil) cartridge in combination with on-line electro- Barrett, E.P., Joyner, L.G., Halenda, P.P., 1951. The determination of
lytic ozonation. Ozone was co-produced by low-voltage pore volume and area distributions in porous substances. I.
electrolysis of water, and the electrolytic ozone could be Computations from Nitrogen isotherms. J. Am. Chem. Soc. 73,
373.
placed either before or after the CDI. The electricity retrieved
Benko, K., Drewes, J.E., 2008. Produced water in the Western
at the discharge of CDI operation could be reused for the United States: geographical distribution, occurrence, and
production of ozone. Water recovery of flow-through capa- composition. Environ. Eng. Sci. in press.
citor increased to more than 90%, and energy recovery rate Bordjiba, T., Mohamedi, M., Dao, L.H., 2007. Synthesis and
was exceeding 30%. Non-charged species including organics electrochemical capacitance of binderless nanocomposite
and pathogens were reduced through ozonation. electrodes formed by dispersion of carbon nanotubes and
carbon aerogels. J. Power Sources 172 (2), 991–998.
Farmer, J.C., Fix, D.V., Mack, G.V., Pekala, R.W., Poco, J.F.,
1996. Capacitive deionization of NH4ClO4 solutions with
4. Conclusions carbon aerogel electrodes. J. Appl. Electrochem. 26,
1007–1018.
 The performance of the CDTs system was consistent Funston, R., Ganesh, R., Leong, L.Y.C., 2002. Evaluation of
throughout the bench-scale experiments and duration of technical and economic feasibility of treating oilfield pro-
duced water to create a ‘‘New’’ water resource. In: Proceedings
the field trials. Deterioration of the carbon-aerogel electro-
of the 2002 GWPC Produced Water Conference, Colorado
des or electrode fouling was not observed throughout the Spring, CO, October 16–17, 2002.
course of the study, but might occur due to adsorption of Gabelich, C.J., Tran, T.D., Suffet, I.H.M., 2002. Electrosorption of
organics during regeneration during times when the inorganic salts from aqueous solution using carbon aerogels.
electrodes are uncharged. Environ. Sci. Technol. 36, 3010–3019.
 The quantity of ions adsorbed to the carbon aerogel (in Hou, C.-H., Liang, C., Yiacoumi, S., Dai, S., Tsouris, C., 2006.
mol/g aerogel) in treating the brackish water was depen- Electrosorption capacitance of nanostructured carbon-based
materials. J. Colloid Interface Sci. 302, 54–61.
dent upon feed water concentrations, increasing in the
IOGCC/ALL Consulting, 2006. A guide to practical management of
order of NabCa4Mg4K for cations, and ClbBr4IClb
produced water from onshore oil and gas operations in the
Br4I for anions. United States. A final report prepared for US Department of
 The maximum percentage of removal, however, followed a Energy, Tulsa, Oklahoma /http://www.all-llc.com/iogcc/
different trend in the order of organic acids (in terms of prodwtr/ProjInfo.htmS.
UVabs at 254 nm) 4I4Br4Ca4alkalinity4Mg4Na4Cl. Matthews, M., 2007. Coalbed methane producers have some new
 Due to the microporous nature of the aerogel structure options. McGraw-Hill Construction/ENR 258, 19.
Nightingale Jr., E.R., 1959. Phenomenological theory of ion
(average pore size of 4.28 nm), the actual available surface
salvation: effective radii of hydrated ions. J. Phys. Chem. 63,
area for ion sorption was about 33% of the measured BET
1381–1387.
area. Therefore, modifying pore size distribution is critical Oh, H.-J., Lee, J.-H., Ahn, H.-J., Joeng, Y., Kim, Y.-J., Chi, C.-S.,
to improve sorption performance of the carbon-aerogel 2006. Nanoporous activated carbon cloth for capacitive
electrodes. deionization of aqueous solution. Thin Solid Films 515 (1),
 CDI could be an alternative for brackish water desalina- 220–225.
tion, but the sorption capacity of the electrodes and Pekala, R.W., Farmer, J.C., Alviso, C.V., Tran, T.D., Mayer, S.T.,
Miller, J.M., Dunn, B., 1998. Carbon aerogels for electrochemi-
operational performance merit further improvement.
cal applications. J. Non-Cryst Solids 225, 74–80.
Ryoo, M.-W., Kim, J.-H., Seo, G., 2003. Role of titania incorporated
on activated carbon cloth for capacitive deionization of NaCl
Acknowledgment
solution. J. Colloid Interface Sci. 264, 414–419.
Shiue, L.-B., Hung, W.-T., Chou, G.-N., Wang, S.-Y., Chung, H.-C.,
The authors thank the US Bureau of Reclamation (BOR) for its Chiu, I.-C., 2005. Energy-effective desalination with online
financial, technical, and administrative assistance in funding sterilization using capacitors. In: Proceedings of the IDA World
and managing the project through which this information Congress, Singapore, September 2005.
ARTICLE IN PRESS
WAT E R R E S E A R C H 42 (2008) 2605 – 2617 2617

Spitz, L.R., 2003. Advanced mobile water treatment systems for Xu, P., Drewes, J.E., Heil, D., 2008. Beneficial use of co-produced
treating coal bed methane produced water. Paper presented at water through membrane treatment: technical–economic
the 2003 Ground Water Protection Council Meeting. assessment. Desalination, in press.
Tran, T.D., Farmer, J.C., Pekala, R.W., 2003. Carbon aerogels and Yang, K.-L., Ying, T.-Y., Yiacoumi, S., Tsouris, C., Vittoratos, E.S.,
their applications in supercapacitors and electrosorption 2001. Electrosorption of ions from aqueous solutions by
processes /http://www.vacets.org/vtic97/tdtran.htmS. carbon aerogel: an electrical double layer model. Langmuir 17,
Veil, J., Puder, M., Elcock, D., 2004. A White paper describing 1961–1969.
produced water from production of crude oil, natural gas, and Yang, C.-Y., Choi, W.-H., Na, B.-K., Cho, B.W., Cho, W.I.,
coal bed methane. A final report prepared for US Department of 2005. Capacitive deionizaton of NaCl solution with carbon
Energy, Argonne National Laboratory, Chicago, Illinois /http:// aerogel-silica gel composite electrodes. Desalination 174,
www.netl.doe.gov/publications/oil_pubs/prodwaterpaper.pdfS. 125–133.
Welgemoed, T.J., Schutte, C.F., 2005. Capacitive deionization Ying, T.-Y., Yang, K.-L., Yiacoumi, S., Tsouris, C., 2002. Electro-
technology: an alternative desalination solution. Desalina- sorption of ions from aqueous solutions by nanostructured
tion 183 (1–3), 327–340. carbon aerogel. J. Colloids Interface Sci. 250, 18–27.
Xu, P., Drewes, J.E., 2006. Viability of nanofiltration and ultra low Zhang, D., Shi, L., Fang, J., Dai, K., Li, X., 2006. Preparation and
pressure reverse osmosis membranes for multi-beneficial use desalination performance of multiwall carbon nanotunes.
of methane produced water. Sep. Purif. Technol. 52, 67–76. Mater. Chem. Phys. 97, 415–419.

You might also like