You are on page 1of 13

Applied Energy 136 (2014) 649–661

Contents lists available at ScienceDirect

Applied Energy
journal homepage: www.elsevier.com/locate/apenergy

On the cost of electrodialysis for the desalination of high salinity feeds


Ronan K. McGovern a,⇑, Adam M. Weiner a, Lige Sun a, Chester G. Chambers a, Syed M. Zubair b,
John H. Lienhard V a
a
Center for Clean Water and Clean Energy, Massachusetts Institute of Technology, Cambridge, MA 02139, USA
b
King Fahd University for Petroleum and Minerals, Dhahran, Saudi Arabia

h i g h l i g h t s

 Feed water was desalinated from 192 down to 0.24 g NaCl/kg solution.
 Energy requirements are similar but treatment costs are lower than evaporation.
 Voltage optimisation can further reduce ED treatment costs by 30–60%.

a r t i c l e i n f o a b s t r a c t

Article history: We propose the use of electrodialysis to desalinate produced waters from shale formations in order to
Received 13 June 2014 facilitate water reuse in subsequent hydraulic fracturing processes. We focus on establishing the energy
Received in revised form 9 September 2014 and equipment size required for the desalination of feed waters containing total dissolved solids of up to
Accepted 15 September 2014
192,000 ppm, and we do this by experimentally replicating the performance of a 10-stage electrodialysis
Available online 11 October 2014
system. We find that energy requirements are similar to current vapour compression desalination pro-
cesses for feedwaters ranging between roughly 40,000-90,000 ppm TDS, but we project water costs to
Keywords:
potentially be lower. We also find that the cost per unit salt removed is significantly lower when removed
Electrodialysis
Desalination
from a high salinity stream as opposed to a low salinity stream, pointing towards the potential of ED to
Brine concentration operate as a partial desalination process for high salinity waters. We then develop a numerical model for
Energy efficiency the system, validate it against experimental results and use this model to minimise salt removal costs by
Hydraulic fracturing optimising the stack voltage. We find that the higher the salinity of the water from which salt is removed
Shale the smaller should be the ratio of the electrical current to its limiting value. We conclude, on the basis of
energy and equipment costs, that electrodialysis processes are potentially feasible for the desalination of
high salinity waters but require further investigation of robustness to fouling under field conditions.
Ó 2014 Elsevier Ltd. All rights reserved.

1. Introduction Water reuse in hydraulic fracturing is of great interest both from an


environmental perspective, as it reduces water use and minimises
We have experimentally investigated factors affecting the cost of disposal through deep-well injection, but also from an economic
electrodialysis (ED) for the desalination of high salinity feeds, focus- perspective as water management costs can account for between
ing on the dependence of the cost of salt removal upon diluate salin- 5% and 15% of drilling costs [1].
ity. We have also developed a numerical model for the system, For the purpose of this investigation, we were most interested
validated it against the experimental results and identified a strat- in flows of water during the life-cycle of a well, which are depicted
egy to optimise the stack voltage such that the sum of equipment in Fig. D.1. For reuse to be economical, the savings in the sourcing,
and energy costs are minimised. Our motivation for this investiga- disposal and transport of water must outweigh any increased costs
tion was the desalination of produced waters in unconventional of treatment or of chemicals in the formulation of the fracturing
oil and gas extraction where, amongst other factors, the presence fluid. This means that regional differences in recycling rates are
of high levels of total dissolved solids can disincentivise water reuse. strongly influenced by regional differences in sourcing, disposal
and transport costs. For example, reuse rates are currently greatest
in the Marcellus shales [3] (reused water makes up 10–15% of the
⇑ Corresponding author. water needed to fracture a well) where transport and disposal
E-mail addresses: mcgov@alum.mit.edu (R.K. McGovern), lienhard@mit.edu (J.H. costs can reach $15–18/bbl ($94–113/m3) [4]. The initial rate at
Lienhard V).

http://dx.doi.org/10.1016/j.apenergy.2014.09.050
0306-2619/Ó 2014 Elsevier Ltd. All rights reserved.
650 R.K. McGovern et al. / Applied Energy 136 (2014) 649–661

Nomenclature

w mass, lbs or kg
Roman symbols x concentration, mol salt/mol water
Am membrane area, m2
C concentration, mol/m3 Greek symbols
D diffusivity, m2/s D change
Es specific energy of salt removal, kW h/lb or kW h/kg  error
Ew specific energy of water produced, kW h/bbl or kW h/m3 K molar conductivity, S m2/mol
h channel height, m l chemical potential, J/mol
i current density, A/m2 m viscosity, m2/s
I current, A Ns specific cost of salt, $/lb or $/kg
k conductivity, S/m Nw specific cost of water, $/bbl or $/m3
KE energy price, $/kW h p osmotic pressure, bar
KQ area normalised equipment price, $/m2 membrane q density, kg/m3
Ls membrane salt permeability, m2/s r spacer shadow factor, –
Lw membrane water permeability, mol/m2 s bar ss specific process time, days/lb or days/kg
m slope sw specific process time, days/bbl or days/m3
M molar mass, kg/mol
ms molal concentration, mol/kg w Subscripts
N number of moles, mol am anion exchange membrane
ncp number of cell pairs, – c concentrate
r membrane surface resistance, X m2 circ circuit
R universal gas constant, J/mol K cm cation exchange membrane
Re Reynolds number d diluate
Sc Schmidt number el electrode
Sh Sherwood number i stage number
t process time, s j time period
T system life, years m membrane surface
T cu integral membrane counterion transport number, – p pump
t cu solution counter-ion transport number, – r rinse
T cp cell pair salt transport number, – s salt
s
T cp cell pair water transport number, – s water
w
V corr stack voltage corrected for concentration polarisation, V
V stack stack voltage, V Superscripts
V volume, m3 f final
_V volume flow rate, m3/s in initial

which produced water flows to the surface (e.g. within the first chloride contents; increased scaling within the shale formation with
10 days) also influences the viability of reuse as low initial pro- the presence of divalent ions; increased corrosion of pipes; increased
duced water volume flow rates making the logistics of reuse more levels of sulphate reducing bacteria resulting in the production of
difficult [3,5]. H2 S gas [8]; and a reduction in the performance of coagulation/floc-
Moving to the costs of reuse, and setting aside the expense asso- culation, flotation, gravity settling and plate and frame dewatering
ciated with logistics, the costs come primarily in the form of: equipment due to residual unbroken polymer gel [9].
increased water treatment costs; increased chemical costs in the Many of the challenges faced in reuse can be dealt with through
formulation of the hydraulic fracturing fluid to mitigate undesir- primary treatment that removes suspended solids, oil, iron, unbro-
able feed water properties; and/or reduced oil or gas production ken polymers and bacteria [9], generally at a cost much below
from the well. By and large, the increase in treatment costs is high- complete desalination (circa $1/bbl ($6.3/m3) compared to $3.50–
est, and the increase in chemical costs lowest, when produced 6.25/bbl ($22–39/m3) for complete desalination [1]). The need for
water is treated with mechanical vapour compression. Vapour the removal of all solids, suspended and dissolved, is less clear.
compression provides high purity water for the formulation of Opinions vary as to the level of total dissolved solids (TDS) that
the hydraulic fracturing fluid but is expensive – ranges of roughly can be tolerated [10] and a complete understanding of issues of
5–8 kW h/bbl (32–50 kW h/m3) of distillate1 [7] and $3.50–6.25/bbl chemical compatibility remains elusive [2]. There is evidence that,
($22–39/m3) of distillate [1] have been reported for the treatment of with improved chemical formulations, high salinity produced
produced waters. Direct reuse, whereby produced water is directly waters may be reused without desalination, particularly in the for-
blended with freshwater before formulation of the fracturing fluid, mulation of fluids for slickwater processes [11–16] (processes with
results, by and large, in the lowest treatment costs but greater chem- high volume flow rates to avoid premature settling of sand, which
ical costs for fluid formulation and perhaps a decline in the well’s serves to maintain fractures propped open) and to some extent for
production. Increased costs associated with reuse, depending on cross-linked gel fracturing processes [17] (lower volume flow rate
the degree of treatment employed, can come in the form of: processes employing low molecular weight guar gum based gels to
increased friction reducer and scale inhibitor demand with high ensure proppant remains suspended). However, the increase in
chemical costs associated with such formulations not evident.
Depending on the fracturing fluid desired, chemical use can be sig-
1
6.4 kW h/bbl (40 kW h/m3) has been reported for 72.5% recovery of feedwater nificant. Fedotov et al. [9] indicated that the use of drag reducing
with total dissolved solids of 50,000 mg/L [6].
R.K. McGovern et al. / Applied Energy 136 (2014) 649–661 651

agents in slickwater fracturing processes, can reach approximately concentrate concentration in each stage to replicate the concen-
1000 ppm (2 lbs per 1000 gallons), while for cross-linked gel frac- tration that would prevail if the concentrate salinity were to be
turing processes chemical use can be much higher and reach determined solely by the rates of salt and water transport across
15,000 ppm (30 lbs/1000 gallons). the membranes (see Appendix A.1). We held the stack voltage
In place of a distillation process, we propose the use of electro- constant at 8 V in all stages and chose this value such that the
dialysis desalination to partially desalt produced water. The objec- current density at the end of the final stage would be 50% of
tive is to achieve a configuration that can reduce water treatment its limiting value (see Appendix A.2).
costs relative to distillation, by avoiding complete desalination, but The experimental apparatus, illustrated in Fig. D.3 involved an
can provide the benefit of reduced total dissolved solids relative to ED200 stack [27] with 17 cell pairs consisting of seventeen Neo-
a direct reuse configuration. At present, a clear illustration of the septa AMX-SB, eighteen CMX-SB membranes, thirty-four 0.5 mm
dependence of ED salt removal costs on feed salinity is not present spacers and two 1 mm end spacers. We employed a GW Instek
in literature, particularly for feed salinities above brackish. A num- GPR-60600 and an Extech 382275 power supply to provide current
ber of studies consider seawater desalination with electrodialysis in the ranges of 0–5 A and 5–20 A respectively. We measured con-
[18], including electrodialysis-reverse osmosis hybrid configura- ductivity on a Jenco 3250 conductivity meter interfacing with
tions [19], but focus upon energy costs alone. Lee et al. [20] con- model 106L (cell constant, K = 1) and model 107N (cell constant,
sider the effect of feed salinity upon the cost of water from a K = 10) probes. We performed experiments in constant voltage
continuous, as opposed to batch, electrodialysis system for brack- mode, with current measured by an Extech EX542 multimeter.
ish feed waters and McGovern et al. [21] analyse the dependence We determined initial diluate and concentrate volumes by sum-
of water costs upon feed and product salinity in their analysis of ming the initial fluid volumes contained within the beakers (1 L
hybrid ED-RO systems for brackish applications. Few studies exist and 3 L for the diluate and concentrate, respectively, in all tests)
that analyse both energy and capital costs for higher salinity feeds with the internal volumes of the diluate and concentrate fluid cir-
[22]. Batch studies of low salinity produced waters report energy cuits (see Appendix A.3). We determined changes in diluate mass
consumption figures of 1.1 kW h/m3 for 90% TDS removal and by tracking the mass of the diluate within the beaker using an
0.36 kW h/m3 for 50% TDS removal from a 3000 ppm TDS stream Ohaus Scout Pro balance with a range of 0–2 kg. Changes in density
[23]. A study at higher feed water salinity reports energy consump- were also accounted for given knowledge of solution conductivities
tion of 12.4 kW h/m3 for 80,000 ppm TDS [24]. A number of exper- versus time.
imental studies, with desalination occurring in a batch mode, To quantify performance we considered certain key metrics.
report the process times required to achieve a final target purity The first metrics are specific process times, based on stage salt
as increasing with the feed salinity [25,26] but leave unclear how removal, ssi , and final stage diluate volume, sw
i :
process times translate into equipment costs. Furthermore, energy
consumption in batch processes is often reported as an average ti
ssi ¼   ð1Þ
kW h/kg salt removed for an entire process without focusing on V in;d in;d
 V fi ;d C fi ;d
i Ci
how this value varies depending upon the diluate, and to a lesser
extent the concentrate, salinity.
ti
In this work, we conduct multiple stages of batch desalination swi ¼ ð2Þ
on an experimental electrodialysis setup such that each stage rep- V fi ;d
licates closely a stage within a continuous process. Furthermore, where t i is the process time for stage i; V in;d and V fi ;d are the initial
i
we relate batch process times and energy consumptions to the pro- and final stage volumes, and C in;d and C f ;d
are the initial and final
i i
duction rate and specific energy consumption that would be stage concentrations. The second metrics are specific energy con-
achieved from an equivalent continuous system. Coupled with a sumption, based on stage salt removal, Esi , and final stage diluate
simple financial model, these metrics allow us to investigate and volume, Ew i :
optimise the dependence of cost upon the feed salinity to a contin-
uous electrodialysis system. P
j Ii;j V i;j Dt i;j
Esi ¼  ð3Þ
2. Methods V i C i  V fi ;d C fi ;d
in;d in;d

P
2.1. Experimental j Ii;j V i;j Dt i;j
Ew
i ¼ ð4Þ
V fi ;d
We performed an experiment to replicate the performance of
a ten stage continuous flow electrodialysis system capable of where Ii;j and V i;j are the stack current and voltage of stage i in time
desalinating a feed stream from 224 mS/cm (195,000 ppm TDS period j of the process. Dt i;j refers to time increment j of the process
NaCl) down to 0.5 mS/cm (240 ppm TDS NaCl). We studied aque- within stage i.

ous NaCl solutions since Naþ and Cl ions account for the vast We used the above performance metrics to compute cost met-
majority of dissolved solids contained within produced water rics, employing the following simplifying assumptions:
samples taken from the Barnett, Eagleford, Fayetteville, Haynes-
ville, Marcellus and Bakken shale plays [1]. Thus, the electrical 1. We set aside pre-treatment, post-treatment, maintenance
conductivity and chemical activity of salts in produced water and replacement costs, focusing solely on the energy cost
samples – both important influencers of the energy consumption and upfront cost of electrodialysis equipment. These costs
and system size in electrodialysis – are well simulated by aque- strongly depend upon feedwater chemistry and can be sig-
ous NaCl solutions of matching total dissolved solids contents. To nificant, e.g. the cost of basic pre-treatment for waters pro-
do this, we ran ten batch experiments, each representing a single duced from shale plays, which might involve basic solids
stage in a continuous process. We chose the diluate conductivi- removal and/or COD and/or BOD reduction, can fall around
ties at the start of each stage such that the diluate conductivity $1/bbl ($6.3/m3) [1].
was halved in each stage and the salt removal was approxi- 2. We neglected pumping power costs (see Appendix B for
mately 50% per stage [20] (see Fig. D.2). We chose the justification).
652 R.K. McGovern et al. / Applied Energy 136 (2014) 649–661

3. We assumed electricity to be priced at K E = $0.15/kW h Salt, water, and charge transport were modelled based upon the
(a conservative estimate of gas powered distributed genera- approach taken in previous work [31,32]. Salt transport was mod-
tion [28]). elled by a combination of migration and diffusion:
4. We assumed equipment costs to scale with membrane area  cp 
Ts i  
and computed these costs by considering an equipment cost J s ¼ Ncp  Ls C s;c;m  C s;d;m ð9Þ
per unit membrane area of K Q ¼ 1500 $/m2 (based upon the
F
capital cost data collected by Sajtar and Bagley [29] and ana- and water transport by a combination of migration (electro-osmo-
lysed by McGovern et al. [21]). This cost data covers plant sis) and osmosis:
sizes up to 40,000 m3 per day and feed salinities up to  
7,000 mg/L. Its extrapolation to higher salinities is based on T cp
wi
 
J w ¼ Ncp þ Lw ps;c;m  ps;d;m ð10Þ
the assumption that similar stack designs would be F
employed at both high and low salinity. In Eq. (10), Ncp is the number of cell pairs, T cp cp
s and T w are the overall
5. We assumed the total installed cost of the equipment to salt and water transport numbers for the cell pair, Ls and Lw are the
equal three times the estimated equipment costs [30]. This overall salt and water permeabilities of the cell pair, C denotes con-
is an estimate since no data exists on commercial scale elec- centration in moles per unit volume, and p denotes osmotic pres-
trodialysis installations in shale plays. sure (calculated employing osmotic coefficients for aqueous NaCl
6. We amortised equipment costs over a twenty-year life, from Robinson and Stokes [33]). The difference between membrane
T ¼ 20 years, assuming an annualised cost of capital of surface concentrations, C s;c;m and C s;d;m , and bulk concentrations, C s;c
r ¼ 10%. and C s;d , was computed via a convection–diffusion based model for
concentration polarisation (see Appendix C.1).
Given these assumptions, we defined the specific cost of salt The stack voltage was represented as the sum of ohmic terms
removal, in $/lb salt (or $/kg salt) and the specific cost of product and membrane potentials:
water $/bbl (or $/m3) from each stage: 
hd hc 2hr
V stack ¼ Ncp r am þ r cm þ þ i þ r cm i þ i
K Q Am rKd C d rKc C c rkr
Nsi ¼ K E Esi þ  s
 T  si ð5Þ
1 1 þ Ncp ðEam þ Ecm Þ þ V el ð11Þ
r
1  1þr
where K is the molar conductivity, itself taken to be a function of
concentration [34,35] and h denotes channel height. k denotes elec-
K Q Am trical conductivity, the subscript r denotes the rinse solution, r
Nwi ¼ K E Ewi þ  w
 T  si ð6Þ
denotes the spacer shadow factor, r denotes the membrane surface
1 1
r
1  1þr resistance of the anion or cation exchange membrane and V el
denotes the sum of the anode and cathode electrode potentials.
where Am is the total membrane area in the stack. Junction potentials associated with concentration differences across
boundary layers were neglected while membrane potentials Eam
2.2. Model and Ecm were computed assuming quasi-equilibrium salt and water
migration through the membranes (see Appendix C.2).
To minimise the cost of salt removal, through optimisation of A series of calibration tests was conducted to establish the val-
the stack voltage, we constructed a semi-empirical model for the ues of T cp cp
s ; T w ; Ls ; Lw ; r m ; r; V el and the Sherwood number Sh
electrodialysis system, which we validated with experimental (see Appendix D). Each test was repeated three times to ensure
results. The process time, energy consumption, and cost of salt repeatability. Bias errors arising from the determination of the dil-
removal from each stage were computed using a numerical model uate circuit volume (Appendix A.3) and of leakage rates from dilu-
that broke each stage into twenty time periods, with an equal ate to concentrate were propagated through the equations defining
change in diluate salinity in each period. During each of these peri- these nine parameters and combined with the random error [Eq.
ods the stack voltage and rates of salt and water transport were (12)] that was determined from the sample standard deviation of
approximated as constant and used, in conjunction with molar results computed from the three tests. Errors were computed at
conservation equations, to determine the conditions at the start a 68% confidence level.
of the next period. Within each stage, the number of moles of salt
and water present in the diluate at the start and the end of each 2tot ¼ 2bias þ 2random ð12Þ
time step j are related to the molar fluxes of salt and water, J s;j Salt and water transport numbers, and T cp T cp
were determined
s w,
and J w;j and the total cell pair area, Am : via constant current migration tests where the diluate and concen-
trate conductivities were close to one another. Salt and water per-
Ns;jþ1  Ns;j ¼ Am J s;j ð7Þ meabilities, Ls and Lw , were determined via diffusion tests with
zero current and initial diluate concentrations close to zero. Mem-
brane resistance, r m , the spacer shadow factor, r, the electrode
Nw;jþ1  Nw;j ¼ Am J w;j ð8Þ
potential, V el , and the Sherwood number, Sh, were determined from
with Njþ1 and N j the number of moles of salt, s, or water, w, at the voltage-current tests at constant diluate and concentrate salinity.
end or start of each time step, Am the total cell pair area of the stack
and J j the average flux across the membrane area of salt, s, or water, 3. Results: Process time, energy consumption and costs
w, at time step j. The concentrate conductivity was approximated as
constant in time for each stage and equal to the value employed in The process time, energy and cost requirements of electrodialy-
experiments. At each instant in time the diluate concentration is sis treatment are shown on a unit salt removal basis in Figs. D.4,
approximated as uniform across the membrane area. This is D.5 and D.6 and on a unit product water basis in Figs. D.7, D.8
because the time taken for fluid to travel the length of the mem- and D.9, in each case illustrating agreement, within error, between
brane (<8 s) is much less than the stage processing time (>120 s the model and the experiment. The deviation between the model
for all stages). and experiment is greatest in the final stages, where the modelled
R.K. McGovern et al. / Applied Energy 136 (2014) 649–661 653

values of energy consumption and process time are highly sensi- Es  V ð18Þ
tive to the electrode potential. This is because the driving force
Thus, while process time varies significantly with stage number
for salt transport is the difference between the stack voltage and
(note the log2 scale in Fig. D.4) specific energy consumption (plot-
the sum of the electrode potentials and membrane potentials
ted on a linear scale in Fig. D.5) remains relatively constant.
(V stack  N cp ðEam þ Ecm Þ  V el ). The sum of membrane potentials,
Given the above explanations for the trends in process time and
Eam þ Ecm , scales with the natural logarithm of the salinity ratio
energy, on the basis of unit salt removal, it is clear that the cost per
(concentrate to diluate) [32] and therefore, in the final stage where
unit of salt removal must remain relatively constant at low stage
the diluate salinity is lowest, the sum of the membrane potentials
numbers (high diluate salinities) but will rise rapidly due to
is greatest – accounting for over 50% of the 8 V applied across the
increasing equipment costs at higher number stages (lower salini-
stack. The remaining voltage driving salt transport is therefore
ties) as seen in Fig. D.6.
highly sensitive to the modelled value of the electrode potential
Combining these insights in Fig. D.4, D.5 and D.6 with the fact that
(V el ¼ 2:13 ± 0.3 V). This sensitivity further posed a difficulty in
salt removal is approximately halved in each stage moving from
modelling desalination, within the final stage, down to 0.5 mS/
stage 3 to stage 10 we can easily explain the trends on a basis of stage
cm (242 ppm TDS). The modelled value of the electrode potential
product water, seen in Fig. D.7, D.8 and D.9. Specific process time on
(in combination with the modelled values of other fitted parame-
the basis of water produced falls with an increasing stage number
ters, see Appendix D) was such that the back diffusion of salt out-
because the processing time per unit of salt removed (Fig. D.4) rises
weighed salt removal via migration before a conductivity of
more slowly than the quantity of salt removed per stage (see
0.5 mS/cm was reached. For this reason, the model of the final
Fig. D.2). Specific energy consumption, on the basis of product water,
stage is for a final diluate conductivity of 0.55 mS/cm rather than
falls because energy consumption per unit of salt removed is
0.5 mS/cm.
approximately constant (see Fig. D.4) and the quantity of salt
The trends in process time, energy and cost are most easily
removed per stage falls rapidly (see Fig. D.2). As a consequence of
explained by considering these quantities on the basis of salt
falling sw and Ew with increasing stage number, the specific cost of
removal. Here we provide scaling estimates that describe the first
water also falls in moving to higher stage numbers, primarily
order variation in process time, energy and cost with stage num-
because the quantity of salt removed per stage is falling rapidly.
ber. The process time s for any given stage scales with the change
in salinity in that stage DS and the inverse of the current I, which
describes the number of moles of salt removed per coulomb of 3.1. Discussion
charge:
DS Not included in the computation of energy in Fig. D.5 or Fig. D.8
s ð13Þ is energy required for pumping, shown in Fig. D.10. These values
I
for pumping power are computed via experimental measurements
Meanwhile, the current scales approximately with the quotient
of the pressure drop across the stack and assuming 100% pump
of the stack voltage over the stack resistance:
efficiency (see Appendix B for detailed calculations). Comparing
V st these values to those for stack energy consumption in Fig. D.8, it
I : ð14Þ
Rst is clear that pumping power only accounts for a significant portion
of total power consumption at low diluate salinity (e.g. stages 10, 9
The stack resistance scales with the sum of the membrane, con-
and 8 where salt removal rates are lowest). Importantly, these val-
centrate and diluate resistances:
ues of pumping power for a laboratory scale system are unlikely to

h h be representative of pumping power consumption in a large scale
Rst  2r m þ þ ð15Þ system. This is because the processing length of the system inves-
rkc rkd
tigated is only 20 cm, meaning that the entrance and exit head
where r is the spacer shadow factor, h is the diluate and concentrate losses have a disproportionately large effect on the pumping power
channel height, k is the solution conductivity of the diluate d or the relative to the frictional pressure drop within the membranes,
concentrate c, and r m is the anion or cation exchange membrane sur- which would be expected to dominate in large scale systems with
face resistance. The process time therefore scales approximately as: larger processing lengths.
 The range of energy requirements for vapour compression
DS h h
s 2r m þ þ ð16Þ shown in Fig. D.8, and of water costs shown in Fig. D.9, correspond
V rkc rkd roughly to a feed salinity equal to that that of the 3th or 4th stages
At high diluate conductivity (lower number stages in Fig. D.2) of electrodialysis. On this basis, Figs. D.8 and D.9 indicate that elec-
the membrane resistance dominates the stack resistance and thus trodialysis can achieve almost complete salt removal with similar
the diluate and concentrate conductivities have a weak effect on energy requirements and lower water costs than vapour compres-
process time. At low diluate conductivity (high number stages) sion [1,6]. Furthermore, considering electrodialysis costs on the
the diluate resistance dominates the stack resistance and the pro- basis of salt removal (Fig. D.6), it is interesting that costs fall signif-
cess time per unit salt removed scales roughly with the inverse of icantly at higher salinity (e.g. in lower number stages). This points
the diluate conductivity. The stack resistance roughly doubles in to the potential of electrodialysis for the partial desalination of
moving from one stage to the next and so too does the specific pro- high salinity feed streams.
cess time. For electrodialysis systems to be realised for high salinity pro-
The energy consumption per unit salt removed, Es , for any given duced waters, further work is required to address the risks of scal-
stage scales with the product of voltage, the current and the pro- ing and fouling. Future work will need to go beyond the simplified
cess time divided by the change in salinity: solution chemistries considered to date [25,26], and in this work.
Produced waters from shale plays have been shown to contain sig-
VIss
Es  ð17Þ nificant concentrations of dissolved solids with low solubility,
DS including silica, iron, barium and calcium [1,36]. Furthermore,
Considering how process time scales in Eq. (13) it is clear that produced waters may contain significant levels of total organic car-
the energy consumption per unit salt removed scales with the quo- bons (up to 160 mg/L, depending on the method used for oil–water
tient of the voltage over the salt transport number: separation [37]), while ED manufacturers advise against feedwater
654 R.K. McGovern et al. / Applied Energy 136 (2014) 649–661

total organic carbon concentrations above 15 mg/L [38] and a num- ments are similar and project that combined equipment and
ber of studies reveal difficulties in removing total organic carbon energy costs are potentially lower for electrodialysis relative to
with traditional filtration methods [39,40]. vapour compression. If partial, as opposed to complete, desalina-
tion of a feed water is required, the prospects for ED are even
4. Voltage optimisation greater as the cost per unit of salt removed is much lower at high
diluate salinities. For example, salt removal from a stream of
Having validated a numerical model for the system we optimise 500 ppm TDS might cost up to four times that of salt removal from
the voltage in each stage to minimise the costs of salt removal. In a stream at 192,000 ppm TDS per unit of salt removed.
Fig. D.11, we compare three distinct strategies that are shown in Beyond our experimental assessment of electrodialysis at high
Fig. D.12: salinities, we have developed and validated a numerical model cov-
ering a range of diluate salinities from 250 ppm up to 192,000 ppm
1. a constant voltage strategy where the voltage is set such that NaCl. This model reveals the importance of optimising the stack
the current density is 80% of its limiting value at the end of voltage to minimise salt removal costs. For the set of equipment
stage 10 (V stack;i ¼16 V, see Fig. D.13); and energy prices examined, we found that brackish water desalina-
2. a constant voltage strategy where the voltage is set such that tion costs are minimised by operating close to the limiting current
the current density is 50% of its limiting value at the end of density, while, for salt removal from higher salinity streams, lower
stage 10 (V stack;i ¼8 V, see Fig. D.13); and stack voltages can allow cost reductions of up to 30%.
3. an optimised strategy where the total costs per stage (equip- This analysis addresses two major considerations affecting the
ment and energy) are numerically minimised using a quadratic viability of ED for the desalination of high salinity produced
method [41] to identify an optimal voltage (V stack;i ). waters, namely the energy and equipment requirements. Given
that ED compares favourably with vapour compression on these
Fig. D.11 reveals that in higher number stages (lower diluate metrics a more detailed analysis of an ED system under field con-
salinities) the strategy of setting the voltage such that the current ditions is warranted. This might include studies of system fouling
is just below its limiting value (e.g., 80%) is a good one as this and scaling when treating more complex feed waters and an anal-
greatly reduces equipment costs. However, at higher salinities ysis of feedwater pre-treatment requirements and costs to ensure
(lower stage numbers), it is best to operate with a lower stack volt- robust operation.
age that allows for reduced energy consumption. Of course,
depending on the relative price of equipment to energy the optimal Acknowledgements
stack voltage for each stage will differ. Higher electricity prices will
drive lower optimal stack voltages and vice versa. Nevertheless, it The authors acknowledge support from the King Fahd Univer-
is clear that the brackish water strategy of setting the current close sity of Petroleum and Minerals through the Center for Clean Water
to its limiting value [20] is not necessarily optimal for the treat- and Clean Energy at MIT and KFUPM under Project Number R15-
ment of higher salinity waters. CW-11. Ronan McGovern and Chester Chambers acknowledge sup-
port from the Hugh Hampton Young Memorial Fellowship and par-
5. Conclusions tial UROP support from the MIT Energy Initiative, respectively.

Our experimental and economic assessment of electrodialysis Appendix A. Determination of experimental conditions
at salinities up to 192,000 ppm NaCl indicates good potential for
the process at high salinities, such as those seen in produced A.1. Determination of the concentrate salinity in each stage
waters from hydraulically fractured shales. For feedwaters with
TDS of roughly 40,000–90,000 ppm, we show that energy require- A key benefit of multi-staging the ED process at high salinities is
the possibility of selecting a different concentrate salinity in each
stage. If the concentrate salinity were to be the same in all stages
Table D.1
Key parameters used to model salt and water transport across membranes in the it would necessarily be significantly greater than the diluate salin-
electrodialysis stack in order to determine steady-state concentrate salinities for each ity in the first stage. This would result in very strong salt diffusion
stage. from concentrate to diluate and water osmosis from diluate to con-
Symbol Value Ref. centrate in the final stages where the diluate salinity would be
much lower than the concentrate. In our experiment we therefore
Membrane performance parameters
Ts 0.97 [42]
chose higher concentrate salinities in stages with higher diluate
Tw 10 [42] salinities and vice versa. In each stage we set the concentrate salin-
Lw 8:12  105 mol/bar m2 s [42] ity equal to the steady state salinity that would be dictated by the
Ls 5:02  108 m/s [42] relative rates of salt and water transport across the membranes:
r am ; r cm 6.0 X cm2 [42]
r 0.69 [31] Js
xs;c ¼ ðA:1Þ
Solution properties Js þ Jw
D 1:61  10 9 2
m /s [33]
tcu 0.5 [43]
where xs;c is the mole fraction of salt in the concentrate at steady
state. To compute each steady-state concentrate value we modelled
Flow properties/geometry
h 0.7 mm –
salt and water transport using the methods of Section 4. Rather than
Am 271 cm2 – modelling the steady state concentrate salinity for each stage we
ncp 17 – approximated its value by considering the molar fluxes of salt and
V circ 0.5367 L – water at the very end of each stage.
Sh 20 [42]
Since the fitted parameters required for the model were not
Operational conditions known a priori, we considered values from the literature for similar
V 8V –
ED experiments (Table D.1). Furthermore, in practice, an ED system
V el 2V [42]
operator may choose to run the stacks with a lower concentrate
R.K. McGovern et al. / Applied Energy 136 (2014) 649–661 655

salinity than could be reached in steady state, perhaps to avoid tive contribution of stack entrance and exit effects to pressure
scale formation. The concentrate salinities chosen for a given appli- drop is large relative to frictional pressure drop through the pas-
cation may not exactly match the present study. Nonetheless, the sages between the membranes.
results obtained remain significant as stack performance is primar-
ily affected by the diluate conductivity and membrane resistance Appendix C. Electrodialysis model
rather than concentrate salinity, as explained in Section 3.
C.1. Concentration polarisation
A.2. Selection of the stack voltage
The difference between bulk and membrane wall concentra-
We selected a constant operating voltage of 8 V, which ensured tions and osmotic pressures is accounted for by a convection–dif-
that we never exceeded 50% of the limiting current density during fusion model of concentration polarisation:
any stage test. We determined the operating voltage from a voltage
 
versus current test performed at the lowest diluate conductivity T cu  t cu i 2h
(0.5 mS/cm), shown in Fig. D.13. DC ¼  ðC:1Þ
D F Sh

A.3. Determination of diluate circuit volume where D is the solute diffusivity, F is Faraday’s constant, h is the
channel height and t cu is the counter-ion transport number in the
We determined the diluate circuit volume by measuring the diluate and concentrate solutions and is approximated as 0.5 for
change in salinity (via conductivity) of the diluate solution follow- both anions and cations. T cu is the integral counter-ion transport
ing the addition of a known amount of salt. number in the membrane that accounts for both migration and dif-
We initially filled the diluate beaker to the 1 L mark with deion- fusion. It is assumed to be equal in the anion and cation exchange
ised water. We then added a small, known mass of salt, ws , to the membranes and approximated as:
beaker and turned the pumps on. We measured the steady-state
conductivity to determine the concentration, C d in mol/L, of the dil- T cp
s þ1
uate circuit:
T cu  : ðC:2Þ
2
kd For the a priori calculations of concentrate salinities in Appendix
Cd ¼ ðA:2Þ
kd A.1, the Sherwood number is computed using the correlation
where kd is the diluate conductivity in S/m and kd is the conduc- obtained by Kuroda et al. [44] for spacer A in their analysis:
tance in m2 =X equiv. We then converted this concentration to Sh ¼ 0:5Re1=2 Sc1=3 ðC:3Þ
molality, ms;d , and solved for the volume of the circuit, Vcirc :
where Sc is the Schmidt number, calculated using the limiting dif-
ws
Vcirc ¼ ðA:3Þ fusivity of NaCl in water [33] and the kinematic viscosity of pure
M s qw ms;d water m [45], both at 25 °C. Re is the Reynolds number defined as:
where Ms is the molar mass of salt (kg/mol) and qw is the density of 2hV
distilled water at 25 °C. After repeating the measurement three Re ¼ ðC:4Þ
m
times, we obtained a diluate circuit volume of 0.54 ± 0.02 L.
where V is the mass averaged velocity in the channel.

Appendix B. Assessment of pumping power C.2. Junction and membrane potentials

We calculated the required pumping power by measuring the Junction potentials associated with concentration polarisation
pressure drop in the diluate circuit, DP, and multiplying by the dil- are neglected (which is compatible with taking the transport num-
uate flow rate, V,_ held at 76 L/h for each stage. To compute the total
ber of both Na and Cl in solution as 0.5), while the sum of the anion
pumping power, we assumed the pressure drops in the diluate and and cation membrane potentials Eam þ Ecm is computed considering
concentrate circuits to be equal and multiplied by a factor of two. quasi-equilibrium migration of salt and water across the
We discounted the pumping power to drive the rinse circuit since membranes:
in a large scale system the number of cell pairs per stack is large
and hence the ratios of diluate and concentrate flow rates to the
T cp
s T cp
rinse flow rate are be small. Eam þ Ecm ¼ ðls;c;m  ls;d;m Þ þ w ðlw;c;m  lw;d;m Þ ðC:5Þ
F F
We made pressure measurements after flushing the stack with
distilled water and neglected the effect of salinity on density and where ls denotes the chemical potential of salt and lw the chemical
viscosity when considering pressure drops in stages with higher potential of water; both calculated employing osmotic coefficients
salinity streams. Multiplying by the specific process time of each and NaCl activity coefficient data from Robinson and Stokes [33].
stage, si , we computed and plotted the specific pumping energy The subscripts c and d denote the concentrate and diluate while
(see Fig. D.10): the subscript m denotes a concentration at the membrane surface.

Ew _ w
p;i ¼ 2V DP si ðB:1Þ
Appendix D. Determination of fitted parameters
For the high salinity stages (numbers 5 and lower), the specific
pumping energy makes up less than 5% of the total specific energy D.1. Sherwood number
consumed and the contribution to the total specific cost of energy
is negligible. In the low salinity stages, the specific pumping energy The Sherwood number was determined via the limiting current
makes up as much as 40% of the total specific energy consumed. density. A current–voltage test was repeated three times for dilu-
However, this number is largely a characteristic of the small pro- ate and concentrate conductivities of 0.5 mS/cm, the results of
cess length of the laboratory scale system used, wherein the rela- which are shown in Fig. D.13. The Sherwood number was then
656 R.K. McGovern et al. / Applied Energy 136 (2014) 649–661

Fig. D.1. Fresh water [1] is mixed with recycled water and chemicals are added, that may include acids, friction reducers, gelling agents and proppant (sand) [2] to form the
hydraulic fracturing fluid. The fluid is then injected into the well at high pressure to create fractures in the underlying shale formation. A portion of this fluid [3], perhaps in
addition to fluid originally contained in the formations, subsequently returns to the surface, at a rate that generally decreases with time, and is known as produced water. The
produced water may be: subjected to levels of treatment that vary from suspended solids removal to complete desalination [1] and recycling; sent to a disposal well; and/or
employed elsewhere as a kill fluid (a fluid used to close off a well after production is complete) or as a salt based drilling fluid [1].

Stage # Stage 1 Stage 2 Stage 3 Stage 4 Stage 5 Stage 6 Stage 7 Stage 8 Stage 9 Stage 10
230 mS/cm 200 mS/cm 190 mS/cm 180 mS/cm 160 mS/cm 150 mS/cm 130 mS/cm 100 mS/cm 77 mS/cm 52 mS/cm
Concentrate (206,000 ppm) (162,000 ppm) (150,000 ppm) (139,000 ppm) (119,000 ppm) (109,000 ppm) (91,600 ppm) (67,200 ppm) (50,000 ppm) (33,000 ppm)

Initial 224 mS/cm 192 mS/cm 128 mS/cm 64 mS/cm 32 mS/cm 16 mS/cm 8 mS/cm 4 mS/cm 2 mS/cm 1 mS/cm
Diluate (195,000 ppm) (152,000 ppm) (90,000 ppm) (40,700 ppm) (19,100 ppm) (9,010 ppm) (4,310 ppm) (2,070 ppm) (1,010 ppm) (492 ppm)

Final 192 mS/cm 128 mS/cm 64 mS/cm 32 mS/cm 16 mS/cm 8 mS/cm 4 mS/cm 2 mS/cm 1 mS/cm 0.5 mS/cm
Diluate (152,000 ppm) (90,000 ppm) (40,700 ppm) (19,100 ppm) (9,010 ppm) (4,310 ppm) (2,070 ppm) (1,010 ppm) (492 ppm) (242 ppm)

Fig. D.2. We designed each of the ten stages such that the diluate conductivity was halved in each successive stage, with the exception of the first two stages. We reduced the
salt removal in the first two stages to avoid the depletion of water in the diluate beaker before the end of a trial. We chose concentrate conductivities based on the rates of salt
and water transport across the membranes (see Appendix A.1).

Conductivity
Heat Exchanger
Probe Concentrate Rinse

Diluate
Diluate
Balance
Concentrate ED Stack Diluate

ED Stack

DC Power
Flow
Pump

Pressure
Valve Electrode
Rinse

Fig. D.3. The electrodialysis setup consisted of a diluate, concentrate, and rinse circuit feeding an ED200 stack. We employed a heat exchanger to regulate the temperature of
the concentrate, with the stack effectively operating as a second heat exchanger to regulate the diluate temperature. We employed valved-rotameters to regulate the flow
rates in each circuit.

determined by considering the following relationship between it with DNaCl the diffusivity of sodium chloride in solution, F Faraday’s
and the limiting current density: constant, C d the diluate concentration, h the concentrate and diluate
channel heights and T s the salt transport number. The Sherwood
DNaCl FC d Sh
ilim ¼  : ðD:1Þ number was found to equal 18 ± 1 (68% confidence).
2h T s2þ1  t cu
R.K. McGovern et al. / Applied Energy 136 (2014) 649–661 657

Fig. D.7. Stage process time per unit of product water.


Fig. D.4. Stage process time per unit of salt removed.

Fig. D.8. Stage energy per unit of product water. The range of energy consumption
for vapour compression is taken from Hayes et al., for 72.5% recovery of feedwater
Fig. D.5. Stage energy consumption per unit of salt removed.
with TDS of 50,000 mg/L (corresponding roughly to stages 4 through 10) [6].

Fig. D.6. Stage cost per unit of salt removed (based on experimental results).
Fig. D.9. Stage cost per unit of product water (based on experimental results). The
range of water costs for vapour compression is taken from Slutz et al., wherein cases
D.2. Spacer shadow factor are considered with feedwater TDS of 49,500 and 80,000 mg/L (corresponding
roughly to stages 4 through 10 and 3 through 10, respectively) [1].
The spacer shadow factor, r, quantifies the conductance of the
diluate and concentrate channels relative to what the conductance (9.7, 19.3 and 29 A/m2) where the voltage-current relationship
would be were there no spacer. When the diluate and concentrate was only weakly affected by concentration polarisation.
solutions are of high conductivity the stack voltage is insensitive to The stack voltage data was first corrected (from V stack to V corr ) to
the spacer shadow factor, since the membrane resistance domi- remove the effects of concentration polarisation, employing the
nates. Therefore, in determining r we considered tests where the Sherwood number from Appendix D.1 and the model described
diluate and concentrate conductivities were low (0.5, 1.5, 2.5 and in Appendix C.1. This allowed the voltage-current relationship to
7.5 mS/cm). We also considered low values of current density be represented by:
658 R.K. McGovern et al. / Applied Energy 136 (2014) 649–661

Fig. D.12. Effect of voltage strategy upon the optimal voltage. At low stage numbers
the V stack;i ¼ 8 V strategy is close to optimal while at high stage numbers the
Fig. D.10. Energy consumption associated with pumping power. V stack;i ¼ 16 V is closest to optimal.

2Nih 2ihr
V corr ¼ ð2ncp þ 1Þirm þ þ þ V el ðD:2Þ
rk rkr
where the terms on the right hand side represent voltage drops
across the membranes, the diluate and concentrate (both at the
same conductivity), the rinse solutions and the electrodes, respec-
tively. Plotting V corr versus the inverse conductivity of the solution
in Fig. D.14 yields r, the slope. Considering:

2Nih
m¼ ðD:3Þ
r
where m is the slope of each line in Fig. D.14, we determined the
spacer shadow effect at the three different current densities. Since
r should be independent of current density we computed its value
as the average of these three values, giving r ¼ 0:64  0:03.
Fig. D.13. Voltage versus current test with diluate and concentrate conductivities of
0.5 mS/cm.
D.3. Electrode potential

At low current densities the electrode potential was computed To determine the electrode potential at higher current densities,
considering the intercept c of each of the lines in Fig. D.14: current–voltage tests were carried out with diluate and concen-
trate conductivities of 25 mS/cm, 75 mS/cm and 150 mS/cm
2ihr
V el ¼ c  ð2N þ 1Þir m þ ðD:4Þ (Fig. D.15). The linearity of these plots at high current densities
rk r (above approximately 240 A/m2) illustrates that neither membrane
At low current densities, the determination of V el is relatively resistance nor electrode potential is a strong function of current
insensitive to the voltage drop across the membranes and the rinse density at high current densities. Furthermore, for these three con-
solutions since both are small. Therefore, even though rm is not ductivities, the range of current densities illustrated is far below
known a priori, it is reasonable to assume r m ¼ 3  104 X m2, in the limiting current density and the voltage correction for concen-
line with the membrane resistance quoted by the manufacturer tration polarisation is thus negligible (i.e. V stack  V corr ). The elec-
[46]. The values of V el found at 9.7, 19.3 and 29 A/m2 were trode potentials, calculated considering the intercept of the linear
2.4 ± 0.1, 1.9 ± 0.3 and 2.2 ± 0.3, respectively. fits shown in Fig. D.15 (see Eq. D.2), for data taken at 25 mS/cm,

Fig. D.11. Effect of voltage strategy upon the cost of salt removal.
R.K. McGovern et al. / Applied Energy 136 (2014) 649–661 659

Fig. D.14. Determination of spacer shadow effect and electrode potential at low
voltage. The markers represent experimental values while the solid lines represent
the fitted equations. Fig. D.17. Water transport number. The markers represent experimental values
while the solid lines represent the fitted equations.

Fig. D.15. Determination of membrane resistances and electrode potentials from


high conductivity data. The markers represent experimental values while the solid Fig. D.18. Salt permeability. The markers represent experimental values while the
lines represent the fitted equations. solid lines represent the fitted equations.

Fig. D.19. Water permeability. The markers represent experimental values while
the solid lines represent the fitted equations.

Fig. D.16. Salt transport number. The markers represent experimental values while
the solid lines represent the fitted equations.
D.4. Membrane resistance

75 mS/cm and 150 mS/cm were found to be 1.5 ± 0.5, 2.4 ± 0.25 At low diluate and concentrate conductivities the stack voltage
and 2.3 ± 0.4 V, respectively. On the basis of electrode potentials is insensitive to the membrane resistance. Thus, we determined
thus being similar at low and high current density, a value of the membrane resistance from the high conductivity data of
V el ¼ 2:13 ± 0.4 V was considered for the model over the entire Fig. D.15. The membrane resistance at each value of conductivity
range of current densities. was determined using the slope of a linear fit,
660 R.K. McGovern et al. / Applied Energy 136 (2014) 649–661

2Nh 2hr Table D.2


m ¼ ð2N þ 1Þrm þ þ ; ðD:5Þ Electrodialysis model parameters.
rk rkr
Symbol Value Ref.
knowing already the value of r from Appendix D.2. The values
of membrane resistance found for solution conductivities of Solution properties
D 1:61  109 m2/s [33]
25 mS/cm, 75 mS/cm and 150 mS/cm were 4:5  104  5  105 ,
tcu 0.5 [43]
2:8  104  3  105 and 3:0  104  5  105 X m2. Thus, the m [45]
8:9  107 m2/s
membrane resistance was modelled as 3:5  104  1  104 X m2
Flow properties/geometry
over the entire range of diluate and concentrate salinities.
h 0.5 mm –
ncp 17 –
D.5. Salt and water transport numbers Sh 18 –
Membrane parameters
Salt and water transport numbers at solution conductivities of r 0.64 ± 0.03 –
7.5 mS/cm, 75 mS/cm, 150 mS/cm and 225 mS/cm were deter- rm 3:5  104  1  104 X m2 –
mined by running tests at constant current and measuring the T cp
s
Eq. (D.10) –
mass of salt and water transported across the membranes in a fixed T cp
w
Eq. (D.11) –
cp
Ls Eq. (D.12) –
amount of time. Three tests were performed at each set of condi-
Lcp
w
Eq. (D.13) –
tions to ensure repeatability. During these tests, an approximately
constant concentrate conductivity was maintained by selecting an Stack parameters
V cp 8V –
initial concentrate solution volume that was three times that of the
V el 2.1 ± 0.4 V –
diluate. The concentrate beaker was filled with NaCl solution of the
desired conductivity and the diluate beaker was filled with NaCl
solution that was 1.5, 5, 15 and 15 mS/cm higher than the concen-
trate conductivity for the 7.5, 25, 75 and 150 mS/cm cases, respec- and mass were recorded versus time. Throughout the tests, the
tively. The pumps were turned on and a constant current was temperature was held constant at 25 °C. The tests were stopped
applied across the stack. The diluate mass and conductivity were after the diluate concentration reached conductivities of 200 lS/
recorded until the diluate conductivity reached a value 1.5, 5, 15 cm, 900 lS/cm, 900 lS/cm and 3200 lS/cm for the four values of
and 15 mS/cm below that of the concentrate for the 7.5, 25, 75 concentrate conductivity respectively. The salt and water perme-
and 150 mS/cm cases, respectively. The salt and mass transport ability coefficients were determined employing Eqs. D.8 and D.9:
numbers were then determined by Eqs. (D.6) and (D.7): Js
Lcp
s ¼ ðD:8Þ
Dws;d F ðC c  D2Cd ÞAm Ncp
T cp
s ¼ ðD:6Þ
ncp IDtM s
Jw
Dww;d F Lcp
w ¼ ðD:9Þ
T cp ¼ : ðD:7Þ DpAm Ncp
w
ncp IDtM w T s
with Am the active membrane area and N cp the number of cell pairs
Here, Dws;d and Dww;d were the changes in the diluate mass for salt in the stack. A second order polynomial fit was applied to the salt
and water respectively, F Faraday’s constant, I the applied current permeabilities and a power-law fit was applied to the water perme-
across the membrane (10 A), N cp the number of cell pairs, and Dt abilities. Bias errors arising from determining the diluate circuit vol-
the process run time. The temperature was held constant at 25 °C ume (Appendix A.3) and leakage were propagated through Eqs. D.8
and the diluate mass was corrected for leakage from diluate to con- and D.9 and combined with the random error [Eq. (12)] arising from
centrate (determined through leakage tests performed at zero cur- the sample standard deviation of results from the three tests run at
rent with deionised water in the concentrate and diluate chambers). the same conditions.
Bias errors arising from determining the diluate circuit volume
(Appendix A.3) and leakage were propagated through Eqs. (D.7) D.7. Summary of model parameters
and (D.6) and combined with the random error [Eq. (12)] that
was determined from the sample standard deviation of results from A summary of the model parameters and equations is provided
the three tests run at the same conditions. As shown in Fig. D.16, the in Table D.2. Membrane salt transport, water transport, salt perme-
salt transport numbers are decreasing with increasing conductivi- ability and water permeability are modelled as:
ties due to the falling charge density of membranes relative to the
6 2
solutions [47]. Fig. D.17 shows that the water transport numbers T cp 5
s ¼ 4  10 Sd þ 4  10 Sd þ 0:96  0:04 ðD:10Þ
are also decreasing with increasing conductivities because of falling
water activity, which reduces the membranes’ capacity to hydrate T cp 5 2 2
w ¼ 4  10 Sc  1:9  10 Sc þ 11:2  0:6 ðD:11Þ
[48].
12 2
Lcp
s ¼ minð2  10 Sd  3  1010 Sd þ 6  108 ;
D.6. Salt and water permeability
2  1012 S2c  3  1010 Sc þ 6  108 Þ  6  109 ½m=s ðD:12Þ
The permeabilities of the membranes to salt and water at solu-
tion conductivities of 7.5 mS/cm, 75 mS/cm 150 mS/cm and Lcp 0:416
 2  105 ½mol=m2 s bar
w ¼ 5Sc ðD:13Þ
225 mS/cm were determined by running tests at zero current with
de-ionised water flowing in the diluate compartment. Three tests
were performed at each value of concentrate conductivity to References
ensure repeatability. During these tests, an approximately constant
concentrate conductivity was maintained by selecting an initial [1] Slutz JA, Anderson JA, Broderick R, Horner PH, et al. Key shale gas water
management strategies: an economic assessment. In: International conference
concentrate solution volume that was three times that of the dilu- on health safety and environment in oil and gas exploration and production,
ate. The pumps were turned on and data for diluate conductivity society of petroleum engineers; 2012.
R.K. McGovern et al. / Applied Energy 136 (2014) 649–661 661

[2] Vidic R, Brantley S, Vandenbossche J, Yoxtheimer D, Abad J. Impact of shale gas [23] Brown T, Frost CD, Hayes TD, Heath LA, Johnson DW, Lopez DA, et al. Produced
development on regional water quality. Science 2013;340(6134). water management and beneficial use. Tech rep. DOE award no: DE-FC26-
[3] Mantell ME. Produced water reuse and recycling challenges and opportunities 05NT15549; 2009.
across major shale plays. In: Proceedings of the technical workshops for the [24] Turek M. Electrodialytic desalination and concentration of coal-mine brine.
hydraulic fracturing study: water resources management. EPA, vol. 600; 2011. Desalination 2004;162:355–9.
p. 49–57. [25] Dallbauman L, Sirivedhin T. Reclamation of produced water for beneficial use.
[4] Shipman S, McConnell D, Mccutchan MP, Seth K, et al. Maximizing flowback Separ Sci Technol 2005;40(1–3):185–200.
reuse and reducing freshwater demand: case studies from the challenging [26] Sirivedhin T, McCue J, Dallbauman L. Reclaiming produced water for beneficial
Marcellus shale. In: SPE eastern regional meeting. Society of Petroleum use: salt removal by electrodialysis. J Membr Sci 2004;243(1):335–43.
Engineers, 2013. [27] PCCell GmbH. ED 200. Lebacher Strasse 60. D-66265 Heusweiler, Germany.
[5] Nicot J-P, Scanlon BR, Reedy RC, Costley RA. Source and fate of hydraulic <http://www.pca-gmbh.com/pccell/ed200.htm>.
fracturing water in the Barnett shale: a historical perspective. Environ Sci [28] Stiles RF, Slezak MS. Strategies for reducing oilfield electric power costs in a
Technol. 2012;46(6):3580–6. deregulated market. SPE Prod Facil 2002;17(03):171–8.
[6] Hayes T, Severin BF, Engineer PSP, Okemos M. Barnett and appalachian shale [29] Sajtar ET, Bagley DM. Electrodialysis reversal: process and cost
water management and reuse technologies. Contract 2012;8122:05. approximations for treating coal-bed methane waters. Desalin Water Treat
[7] Horner P, Slutz JA. Shale gas water treatment value chain – a review of 2009;2(1–3):284–94.
technologies, including case studies. In: SPE annual technical conference and [30] Peters MS, Timmerhaus KD, West RE, Timmerhaus K, West R. Plant design and
exhibition, no. SPE 147264. Society of Petroleum Engineers; 2011. economics for chemical engineers, vol. 4. New York: McGraw-Hill; 1968.
[8] Proceedings and minutes of the hydraulic fracturing expert panel XTO [31] Fidaleo M, Moresi M. Optimal strategy to model the electrodialytic recovery of
facilities. Fort Worth; September 26th, 2007. a strong electrolyte. J Membr Sci 2005;260(1):90–111.
[9] Fedotov V, Gallo D, Hagemeijer PM, Kuijvenhoven C, et al. Water management [32] McGovern RK, Zubair SM, Lienhard V JH. The cost effectiveness of
approach for shale operations in north america. In: SPE unconventional electrodialysis for diverse salinity applications. Desalination 2014;348:57–65.
resources conference and exhibition-asia pacific. Society of Petroleum [33] Robinson R, Stokes R. Electrolyte solutions. Courier Dover Publications; 2002.
Engineers; 2013. [34] Shedlovsky T. The electrolytic conductivity of some uni-univalent electrolytes
[10] Rassenfoss S. From flowback to fracturing: water recycling grows in the in water at 25C. J Am Chem Soc 1932;54(4):1411–28.
Marcellus shale. J Petrol Technol 2011;63(7):48–51. [35] Chambers J, Stokes JM, Stokes R. Conductances of concentrated aqueous
[11] Bryant J, Robb I, Welton T, Haggstrom J. Maximizing friction reduction sodium and potassium chloride solutions at 25C. J Phys Chem
performance using flow back water and produced water for waterfrac 1956;60(7):985–6.
applications. In: AIPG Marcellus shale hydraulic fracturing conference, 2010. [36] Thiel GP, Lienhard V JH. Treating produced water from hydraulic fracturing:
[12] Kamel A, Shah SN. Effects of salinity and temperature on drag reduction composition effects on scale formation and desalination system selection.
characteristics of polymers in straight circular pipes. J Petrol Sci Eng Desalination 2014;346:54–69.
2009;67(1):23–33. [37] Silva JM. Produced water pretreatment for water recovery and salt production.
[13] Paktinat J, Bill O, Tulissi M. Case studies: improved performance of high brine Tech rep. Research Partnership to Secure Energy for America; 2012.
friction reducers in fracturing shale reservoirs. Soc Petrol Eng 2011:1–12. [38] GE Power & Water. GE 2020 EDR Systems; 2013. <https://www.gewater.com/
[14] Aften CW. Study of friction reducers for recycled stimulation fluids in kcpguest/documents/Fact%20Sheets_Cust/Americas/English/FS1242EN.pdf>.
environmentally sensitive regions. In: SPE eastern regional meeting. Society [39] Jiang Q, Rentschler J, Perrone R, Liu K. Application of ceramic membrane and
of Petroleum Engineers; 2010. ion-exchange for the treatment of the flowback water from Marcellus shale
[15] Zhou MJ, Baltazar M, Qu Q, Sun H. Water-based environmentally preferred gas production. J Membr Sci 2013;431:55–61.
friction reducer in ultrahigh-TDS produced water for slickwater fracturing in [40] Michel MM, Reczek L. Pre-treatment of flowback water to desalination.
shale reservoirs. In: SPE/EAGE European unconventional resources conference Monogr Environ Eng Committee 2014;119:309–21.
and exhibition. Society of Petroleum Engineers; 2014. [41] Klein SA. Engineering equation solver. Academic professional V9.438-3D;
[16] Hallock JK, Roell RL, Eichelberger PB, Qiu XV, Anderson CC, Ferguson ML. 2013.
Innovative friction reducer provides improved performance and greater [42] Fidaleo M, Moresi M. Electrodialytic desalting of model concentrated nacl
flexibility in recycling highly mineralized produced Brines. In: SPE Brines as such or enriched with a non-electrolyte osmotic component. J
unconventional resources conference-USA. Society of Petroleum Engineers; Membr Sci 2011;367(1):220–32.
2013. [43] Sonin A, Probstein R. A hydrodynamic theory of desalination by electrodialysis.
[17] LeBas R, Lord P, Luna D, Shahan T. Development and use of high-TDS recycled Desalination 1968;5(3):293–329.
produced water for crosslinked-gel-based hydraulic fracturing. In: 2013 SPE [44] Kuroda O, Takahashi S, Nomura M. Characteristics of flow and mass transfer
hydraulic fracturing technology conference; 2013. rate in an electrodialyzer compartment including spacer. Desalination
[18] Turek M. Dual-purpose desalination-salt production electrodialysis. 1983;46(1):225–32.
Desalination 2003;153(1):377–81. [45] Association ER. Viscosity of water and steam. Edward Arnold Publishers; 1967.
[19] Galama A, Saakes M, Bruning H, Rijnaarts H, Post J. Seawater predesalination [46] ASTOM Corporation. Neosepta; 2013. <http://www.astom-corp.jp/en/en-
with electrodialysis. Desalination 2013;342:61–9. main2-neosepta.html>.
[20] Lee H-J, Sarfert F, Strathmann H, Moon S-H. Designing of an electrodialysis [47] Kontturi K, Murtomaki L, Manzanares JA. Ionic transport processes in
desalination plant. Desalination 2002;142(3):267–86. electrochemistry and membrane science. Oxford; 2008.
[21] McGovern RK, Zubair SM, Lienhard V JH. The benefits of hybridising [48] Berezina N, Gnusin N, Dyomina O, Timofeyev S. Water electrotransport in
electrodialysis with reverse osmosis. J Membr Sci 2014;469:326–35. membrane systems. Experiment and model description. J Membr Sci
[22] McGovern RK, Zubair SM, Lienhard V JH. Hybrid electrodialysis reverse 1994;86(3):207–29.
osmosis system design and its optimization for treatment of highly saline
Brines. IDA J Desalination Water Reuse 2014;6(1):15–23.

You might also like