Sei sulla pagina 1di 8

Electrochimica Acta 107 (2013) 611618

Contents lists available at SciVerse ScienceDirect

Electrochimica Acta
journal homepage: www.elsevier.com/locate/electacta

Efcient extraction of lignin from black liquor via a novel


membrane-assisted electrochemical approach
Wei Jin a , Rasha Tolba a , Jiali Wen a , Kecheng Li b , Aicheng Chen a,,1
a
b

Department of Chemistry, Lakehead University, 955 Oliver Road, Thunder Bay, ON P7B 5E1, Canada
Department of Chemical Engineering, University of New Brunswick, 15 Dineen Drive, Fredericton, NB E3B 5A3, Canada

a r t i c l e

i n f o

Article history:
Received 26 April 2013
Received in revised form 6 June 2013
Accepted 6 June 2013
Available online xxx
Keywords:
Black liquor
Lignin extraction
Water electrolysis
Ion exchange
Mass transport

a b s t r a c t
A novel membrane-assisted electrochemical approach for the precipitation of lignin from black liquor
has been developed. Without the addition of acid or CO2 , the pH in the black liquor solution was lowered
to 4.7 due to water electrolysis, leading to pH-dependent lignin precipitation. Simultaneously, Na+ ions
traversed the membrane to the cathode compartment to balance the OH that was generated in the
cathodic reaction, facilitating caustic recovery. Owing to lignin precipitation and oxidation, greater than
70% of the chemical oxygen demand (COD) was removed, which surpasses the efcacy of COD reduction
via conventional acidic precipitation. In addition, it has been demonstrated that the pH change was signicantly inuenced by the electrolyte condition in the system. With decreasing cation concentration in
the cathode compartment, Na+ transport through the membrane was facilitated, which inhibited competitive H+ transport and improved the performance of the electrochemical cell. This novel approach may
serve as a promising alternative for the cost-effective extraction of lignin and recovery of NaOH from the
black liquor with modulation feasibility.
2013 Elsevier Ltd. All rights reserved.

1. Introduction
Black liquor is a highly viscous aqueous waste that is generated by the alkaline Kraft process in wood pulping, consisting of
inorganic pulping chemicals as well as organics from extracted
wood constituents [1]. The solid content of black liquor varies
between 15% and 40% by weight, whereas lignin makes up 3045%
of the total solid composition [2,3]. Lignin is an amorphous complex
polyphenolic plant constituent that is the second most abundant
plant derived polymer (after cellulose) [4]. Although million tons
of lignin is produced every year in black liquor, only a minimal
amount is separated through direct extraction from plants [5]. In
the Kraft process, over 90% of lignin is simply burned to recover
the chemicals from black liquor. Extract some lignin from black
liquor would reduce the load of the Kraft recovery system, and it
thus to increase pulping capacity. On the other hand, the presence
of lignin in wastewater streams signicantly increases the chemical oxygen demand (COD) and biological oxygen demand (BOD),
which are responsible for serious damage to the environment and
human health [6]. Conversely, there is strong potential for deriving great economic value from the lignin in black liquor, which

Corresponding author. Tel.: +1 807 343 8318; fax: +1 807 346 7775.
E-mail address: aicheng.chen@lakeheadu.ca (A. Chen).
1
ISE member.
0013-4686/$ see front matter 2013 Elsevier Ltd. All rights reserved.
http://dx.doi.org/10.1016/j.electacta.2013.06.031

may be employed as raw material for the production of vanillin,


phenols, activated carbons, etc., or be converted to combustible
fuel gases such as hydrogen, methane and CO via gasication
[7].
Due to the great importance of lignin in environmental and
economical consideration, substantial efforts have been made for
its extraction from black liquor using various biological, chemical and electrochemical techniques [810]. Biological technologies
involve the use of fungi and bacteria through aerobic and anaerobic schemes [10], although the efciency of the microorganisms is
limited by the presence of toxic compounds within the black liquor.
Supercritical uid extraction and solvent extraction have also been
attempted in the removal of low molecular weight lignin from
black liquor; however, these techniques suffer from high operating
costs [11]. Recently, economically attractive methods have been
achieved by acid precipitation (particularly sulfuric acid) and CO2
precipitation owing to their capacities for color and COD reduction
as well as cost effectiveness [12]. These processes have developed
due to the fact that pH is the dominating parameter that controls the
solubility of Kraft lignin, where lignin proceeds to precipitate when
the pH of the black liquor solution is below 10.5 [1315]. However,
the highly acidic ltrate and the associated operation problems
limit the wide application of acid precipitation [16]. Meanwhile,
although it has been identied that lignin precipitation from black
liquor via CO2 acidication is superior to acid precipitation as it
has lower operating costs and provides a precipitate that is more

612

W. Jin et al. / Electrochimica Acta 107 (2013) 611618

easily ltered, its chief drawback is that the pH cannot be reduced


any lower than 8 [17]. In this regard, acid precipitation can produce higher lignin yields than its CO2 counterpart, as it may attain
a much lower pH.
Electrochemical techniques have been employed for organics
separation from black liquor, due to the fact that it can generate sufcient occulant in situ to reduce the chemicals that are employed
and subsequently disposed of [18,19]. Electrolysis treatment was
carried out on black liquor at different electrode potentials. Complete COD removal was achieved at a higher potential treatment
with the oxidation of lignin at the anode and hydrogen production
at the cathode; however, it required signicantly higher energies spanning from 17,700 J at 1.5 V to 149,000 J at 12 V/g of COD
removal [20]. Furthermore, membrane-assisted electrolysis (also
called electrodialysis) was employed as a pretreatment of black
liquor [21]. However, the majority of these investigations focused
on the removal of COD and caustic recovery, rather than lignin
extraction [21,22].
As mentioned above, the solubility of lignin in black liquor can be
determined to a signicant degree via its pH, providing a potential
solution for improving membrane-assisted electrochemical processes. It has been suggested that the pH might be modulated by
electrochemical reactions in conjunction with a cationic exchange
membrane, leading to the caustic recovery and lignin precipitation
from black liquor. In this regard, the aim of this study is to develop
a membrane-assisted electrochemical method for the treatment of
black liquor, wherein a Naon membrane was employed due to
its excellent operation stability [23]. This approach may serve as a
promising alternative for overcoming the disadvantages of existing
conventional black liquor treatments.

2. Experimental
2.1. Chemicals
The chemicals utilized in this study included iridium chloride hydrate IrCl3 3H2 O (Pressure Co.), tetrabutyl orthotitanate
(Fluka), isopropyl alcohol (SigmaAldrich), tantalum pentachloride Ta2 Cl5 (NOAH Technologies Corporation), sodium hydroxide
(Anachemia), sodium sulfate (Fisher Scientic), sulfuric acid
(Aldrich), and diethyl ether (Aldrich), all of which were analytical
grade chemicals. Acetonitrile (Caledon Laboratory, Ltd.) was HPLC
grade. Pure water (18 M cm) was obtained from a NANO pure
DiamondTM UV ultrapure water purication system. The lignin
sample used for comparison was prepared from black liquor by
passing carbon dioxide (CO2 ) through a black liquor solution that
was provided by a local pulp and paper mill. The precipitated

lignin was subsequently puried by rinsing with a dilute sulphuric


acid solution. The solutions used in this research were prepared
by dissolving the corresponding chemicals in ultrapure water.
2.2. Fabrication and characterization of Ti/IrO2 -Ta2 O5 electrodes
The Ti/TiO2 IrO2 electrode was prepared by a thermal decomposition technique [24]. The Ti substrates strips were cleaned in
an ultrasonic bath of acetone for 10 min followed by 10 min in distilled water, and then etched in 18% HCl at approximately 85 C for
30 min. Meanwhile, the coating solution was prepared by mixing
the tantalum precursor (0.14 g Ta2 Cl5 dissolved in 7.5 ml isopropanol) and the iridium precursor (0.30 g IrCl3 3H2 O dissolved in
2.5 ml of ethanol). Subsequently, the coating solution was painted
onto the Ti substrates using the brush technique, and was repeated
until a 30 g/m2 oxide coating load was obtained. Finally, the electrodes were annealed at 450 C for 1 h. The surface morphology and
composition of the coatings were characterized by a JEOL 5900LV
scanning electron microscope (SEM), which was equipped with
X-ray energy dispersive spectrometry (EDS).
2.3. Electrochemical investigations of lignin solution and black
liquor
A VoltaLab (PGZ-301 Dynamic-EIS Radiometer) potentiostat
was employed for the present electrochemical investigation. A Ushaped electrochemical cell as shown in Fig. 1 was segmented
into the anode and cathode compartments, which were separated by a Naon N117 membrane (4.5 cm in diameter 183 m
thick, DuPont Company). The fabricated Ti/IrO2 Ta2 O5 electrode
(2.0 cm2 ) was employed as the working electrode, and the reference electrode was Hg/HgO electrode. A Pt mesh (10 cm2 ) counter
electrode was cleaned prior to each experiment via ame annealing
and quenched with pure water. The electrolytes for electrochemical study were deaerated with ultrapure argon gas prior to
measurements. Cyclic voltammograms (CVs) for the different electrolyte systems were recorded out in the electrode potential region
between 0.80 and 1.45 V at a scan rate of 50 mV/s. Electrochemical impedance spectra of mass transport effect were measured from
100 kHz to 7.5 mHz with an a.c. voltage amplitude of 10 mV. For a
comparison to the electrochemical treatment, the acidication of
the 10% 50 ml black liquor solution was carried out by the dropwise
addition of a 4 M sulfuric acid solution with rapid stirring. The pH
was monitored until a pH of 4.7 was attained.
The durability and permeability properties of the Naon
membrane were measured by monitoring its proton conductivity
after several cycles of the consecutive electrochemical treatment

Fig. 1. Schematic diagram of the experimental setup designed for the membrane-assisted electrochemical treatment of black liquor.

W. Jin et al. / Electrochimica Acta 107 (2013) 611618

of the black liquor solution. Prior to each conductivity measurement, the membrane was rinsed with deionized water and then
sandwiched between two platinum plate electrodes supported
Teon plates. Subsequently, this cell was placed in deionized water
and impedance spectrum was recorded from 100 Hz to 10 mHz
with an amplitude of 10 mV. The conductivity () was calculated
using the following equation:
=

L
R A

(1)

where L is the distance between the two electrodes; A is the geometric area of the membrane; and R is the membrane impedance
value derived from the low intersect of the high frequency
semicircle.
2.4. Instrumental analysis
The lignin precipitated from the electrochemical treatment and
acidication of the 10% black liquor solution was separated by
centrifugation, respectively. The COD test was carried out using
174-318 accu-TEST standard range (20900 mg/l) twist cap vials. A
2.5 ml centrifugate solution was transferred to a vial that contained
a chromic acid solution, heated to 150 C for two hours and then
allowed to cool. The UV absorbance was recorded at 600 nm using a
HACH-DR 2800 portable spectrophotometer. Further, an additional
1 ml centrifugate solution was extracted three times with 5 ml of
diethyl ether, respectively. The resulting product was analyzed
using HPLC (Varian Prostar 230 with a Symmetry C8 column containing dimethyloctylsilyl bonded amorphous silica-acetonitrile).
In order to determine the caustic recovery, the sodium ion concentrations of the cathode and anode compartments were measured
using inductively coupled plasma (ICP, Perkin-Elmer Optima 5300
DV).

613

3. Results and discussion


3.1. Characterization of the fabricated Ti/Ta2 O5 IrO2 electrode
As shown in Fig. 2A, a SEM image of the Ta2 O5 IrO2 coating
exhibits a cracked-mud structure, which is typical for oxide electrodes due to the sintering process [25]. On the electrode surface,
some crystalline agglomerates are scattered within at areas or
along ridges, as shown in Fig. 2B and C, which is consistent with
previously reported Ta2 O5 IrO2 coated Ti electrodes [26,27]. Additionally, strong Ir, Ta and oxygen peaks are observed in the EDS
spectrum of the Ta2 O5 IrO2 coating (Fig. 2D), while a small peak of
Ti is derived from the corresponding Ti substrate. Further quantitative analysis conrmed that 30% Ta2 O5 is present in the coating,
which is consistent with the coating precursor composition.
A U-shaped electrochemical cell was designed and divided into
two compartments, which were separated by a Naon membrane,
as illustrated in Fig. 1. This membrane plays a signicant role as
it selectively allows cations (sodium ions or H+ ) to freely pass
through it. The Ti/Ta2 O5 IrO2 electrode was chosen as the anode
since our previous studies had shown that the Ti/Ta2 O5 IrO2 electrode possesses high catalytic activity toward oxygen evolution and
high stability [24,28]. Fig. 3 presents the CVs of the Ti/Ta2 O5 IrO2
electrode recorded in the blank, lignin and black liquor solutions,
respectively. The CV curves for the blank and lignin solutions have
similar shapes, showing the onset potential of oxygen evolution at
nearly 0.56 V, the redox couple reaction of IrO3 /IrO2 at 0.10 V (oxidation) and 0.35 V (reduction). However, as compared to the CV
curve in the blank solution, the current density of oxygen evolution
decreases slightly in the lignin solution, which is likely due to the
adsorption of lignin and its oxidation intermediate at the electrode
surface [24,28]. In contrast, the current density of oxygen evolution
signicantly decreases and its onset potential positively shifts to

Fig. 2. SEM images of the Ti/Ta2 O5 -IrO2 electrode (A) general view, (B) at area, (C) ridge part, and the corresponding (D) EDS spectrum.

614

W. Jin et al. / Electrochimica Acta 107 (2013) 611618

40
-2

Current density / mAcm

anodic reaction proceeds (Eq. (2)) in an alkaline solution, while it


transfers to (Eq. (3)) in an acidic solution, resulting in a sharp drop
in pH.

a
b
c

30

2H2 O O2 + 4H+ + 4e

20
10
0

-10
-1.0

-0.5

0.0

0.5

1.0

1.5

Potential vs. (SCE) / V


Fig. 3. Cyclic voltammetric curves of the Ti/Ta2 O5 -IrO2 electrode recorded in the different membrane systems at a scan rate of 50 mV/s. Cathode compartment: 0.05 M
NaOH + 0.20 M Na2 SO4 ; anode compartment: (a) 0.05 M NaOH + 0.20 M Na2 SO4 ; (b)
100 ppm lignin + 0.05 M NaOH + 0.20 M Na2 SO4 ; and (c) 10% black liquor solution.

0.79 V in the black liquor solution, indicating the inhibited effect


of its complex composition toward electrochemical treatment. This
fact may also be indicated from the disappearance of the IrO3 /IrO2
redox couple reaction in the CV curve for the black liquor solution.
Consequently, it is suggested that the pH of the black liquor solution might be decreased due to the electrochemical reaction (Eq.
(2)), and the composition of the system plays an important role in
the performance of the cell.
4OH O2 + 2H2 O + 4e

(2)

3.2. pH change in lignin solution system


The electrochemical performance of a Ti/Ta2 O5 IrO2 electrode
in a lignin solution was characterized using chronopotentiometry
at a 100 mA/cm2 current density. As shown in Fig. 4A, the pH of
the anode side (lignin solution) decreased from 12.7 to 2.2 within
45 min (curve a), while a sharp reduction was observed, from 10.0
at 30 min to 4.0 at 35 min. It has been demonstrated that the electrochemical reaction is highly pH-dependent [29,30] and that this
sharp reduction may possibly be due to the fact that there is a reaction pathway transition occurring at the anode [31]. Initially, the

12

a
b
c

Accordingly, as is depicted in Fig. 4B (curve a), the pH at the cathode compartment increases from 12.7 to 13.2 due to the reaction
of (Eq. (4)). However, the slope for pH change gradually decreases
with increasing reaction time, suggesting competitive cation transport between Na+ and H+ . It has been demonstrated that the cation
transport is largely determined by the electrolyte condition within
the membrane system [32,33]. With the alteration of the Na+ concentration gradient between the anode and cathode compartments,
the membrane resistance for Na+ transport to the cathode compartment greatly increases. In order to balance the negative charge of
OH that is produced by (Eq. (4)), there is increasing H+ transport
to the cathode compartment, which leads to the pH slope reduction. It is suggested that the change in pH might be modulated by
the modication of the electrolyte condition with respect to the
cation transport. Consequently, when the initial Na+ concentration
in the cathode compartment decreases from 0.45 M to 0.13 M, the
pH decrease rate signicantly improved in the anode compartment
as shown in Fig. 4A (curve b), and the pH increases linearly with a
larger slope, as shown in Fig. 4B (curve b).
2H2 O + 2e 2OH + H2

pH

(4)

The change in pH may also be accelerated by increases the current density. As illustrated in Fig. 4A (curve c) and B (curve c), the
corresponding pH changes at the anode and cathode compartments
are facilitated at a current density of 150 mA/cm2 as compared to
those at 100 mA/cm2 . These facts may also be suggested in the
chronopotentiometry curves as shown in Fig. 5. With the increasing reaction time, the electrode potential initially linearly increased
due to the electrolyte pH change in the two compartments of membrane system. The potential change rate increases via a reduction
in the Na+ concentration at the cathode compartment, or by an
increase in the current density, consistent with the pH trend shown
in Fig. 4. Interestingly, the potential begins to decrease at the exact
point in time where the anodic pH drops to below 7, which further
conrms electrolyte-dependent effects on cell performance.
3.3. Modulated electrochemical treatment of black liquor solution
In order to illustrate the effect of mass transport, the electrochemical impedance spectroscopy (EIS) of different systems was

12.8

(3)

a
b
c

pH

12.6

12.4

3
0

10

20

30

t / min

40

50

60

12.2

10

20

30

40

50

60

t / min

Fig. 4. pH changes of (A) anode and (B) cathode during the electrochemical treatment of the lignin solution. Anode compartment: 100 ppm lignin + 0.05 M NaOH + 0.20 M
Na2 SO4 ; cathode compartment: (a) 0.05 M NaOH + 0.20 M Na2 SO4 at 100 mA/cm2 current density; (b) 0.05 M NaOH + 0.04 M Na2 SO4 at 100 mA/cm2 current density; and (c)
0.05 M NaOH + 0.04 M Na2 SO4 at 150 mA/cm2 current density.

W. Jin et al. / Electrochimica Acta 107 (2013) 611618

Potential vs. (SCE) / V

3.0

a
b
c

2.7
2.4
2.1
1.8
1.5
0

10

20
30
t / min

40

50

Fig. 5. Chronopotentiometric curves during the electrochemical treatment of


the lignin solution. Anode compartment: 100 ppm lignin + 0.05 M NaOH + 0.20 M
Na2 SO4 ; cathode compartment: (a) 0.05 M NaOH + 0.20 M Na2 SO4 at 100 mA/cm2
current density; (b) 0.05 M NaOH + 0.04 M Na2 SO4 at 100 mA/cm2 current density;
and (c) 0.05 M NaOH + 0.04 M Na2 SO4 at 150 mA/cm2 current density.

20

investigated. According to electrochemical theory, the impedance


of the high frequency region is determined by charge transfer
or heterogeneous transport, while the impedance of the low frequency region is dependent on mass transport or species diffusion
[34]. As shown in Fig. 6A, Nyquist plots exhibit two well-dened
semicircles. The smaller semicircles residing in the high frequency
region for different systems are nearly identical, which might be
attributed to the impedance of the oxide electrode coating [24]. It
appears obvious that the mass transport impedance at a low frequency is dependent on the condition of the electrolyte. As is shown
in Fig. 6A (curve a and c), the mass transport of a 100 ppm lignin
solution increases signicantly as compared to a corresponding
blank solution, while this impedance can be reduced by decreasing
the Na+ concentration at the cathode compartment, as revealed in
Fig. 6A (curve b), consistent with the results mentioned above.
The mass transport impedance of the black liquor solution system (curve f of Fig. 6B) is signicantly increased as compared to the
corresponding lignin solution system (curve c of Fig. 6A), which is in
agreement with the conclusion suggested in Fig. 3. Similarly, mass
transport impedance may be reduced by decreasing the Na+ concentration at the cathode compartment as shown in Fig. 6B (curve
d and e). It should be noted that the electrolyte utilized at the cathode compartment is a mixture of NaOH and Na2 SO4 solution, where
the NaOH solution provides the same initial pH as the anode compartment. The Na2 SO4 solution provides additional ionic species

a
b
c

90

12

d
e
f

60

30

4
0

-Zim /

-Zim /

16

615

0
0

10

20

30

Zre /

40

50

40

80

120

Zre /

160

200

Fig. 6. EIS curves of the lignin and black liquor solutions recorded at a potential of 750 mV. (a) 0.05 M NaOH + 0.20 M Na2 SO4 solution in anode and cathode compartments, respectively; (b) 100 ppm lignin in 0.05 M NaOH + 0.20 M Na2 SO4 solution (anode compartment), 0.05 M NaOH + 0.04 M Na2 SO4 solution (cathode compartment); (c)
100 ppm lignin in 0.05 M NaOH + 0.20 M Na2 SO4 solution (anode compartment), 0.05 M NaOH + 0.20 M Na2 SO4 solution (cathode compartment); (d) 10% black liquor solution
(anode compartment), 0.05 M NaOH solution (cathode compartment); (e) 10% black liquor solution (anode compartment), 0.05 M NaOH + 0.04 M Na2 SO4 solution (cathode
compartment); and (f) 10% black liquor solution (anode compartment), 0.05 M NaOH + 0.20 M Na2 SO4 solution (cathode compartment).

14

a
b
c

12

13.8
13.5

pH

10

pH

a
b
c

13.2

8
6

12.9

4
0

50

100

150

t / min

200

250

12.6

50

100

150

200

250

t / min

Fig. 7. pH changes of (A) anode and (B) cathode compartments during the electrochemical treatment of 10% black liquor solution (anode compartment), cathode compartment:
(a) 0.05 M NaOH + 0.20 M Na2 SO4 at 100 mA/cm2 ; (b) 0.05 M NaOH + 0.04 M Na2 SO4 at 150 mA/cm2 ; (c) 0.05 M NaOH at 150 mA/cm2 .

616

W. Jin et al. / Electrochimica Acta 107 (2013) 611618

Table 1
Comparison of the performance for 10% black liquor solution treatment using different operation methods and conditions.
Treatment

Current density (mA/cm2 )

Cathodic solution

t (min)

COD (mg/l)

Precipitated lignin (g)

Na+ recovery

Final pH

EC
EC
EC
Acid

100
150
150

0.05 M NaOH + 0.20 M Na2 SO4


0.05 M NaOH + 0.04 M Na2 SO4
0.05 M NaOH

220
80
75

4781
5310
5520
5975

1.01
1.06
1.09
1.45

51%
43%
49%

4.72
4.70
4.69
4.70

3.4. Lignin extraction from black liquor solution


Different procedures for the precipitation of lignin from
black liquor were studied, including acidication and membraneassisted electrochemical treatments, while their nal pH values
were virtually the same as shown in Table 1. The removal of organic
compounds (primarily lignin) from black liquor solutions, which
is of vital importance for pulp mill wastewater treatment, can be
estimated by determining COD values. The COD decreased with
increasing EC treatment time, and more than 70% of the initial COD
value was reduced (Fig. 8A and Table 1), leading to a signicant
reduction in the color of the solution (Fig. 8B). With alternate treatment methods, the nal COD value was variable. The removal of
COD via electrochemical treatment was superior to that of acidication, where longer duration electrochemical treatment translates
to improved COD removal. This may attributed to the generation of
a number of value-added intermediates, such as vanillin and vanillic acid, which were generated via electrochemical lignin oxidation

60

a
b
c

50

Na recovery/%

toward the improvement of the total conductivity, which will be


discussed further below.
The electrochemical treatment of a 10% black liquor solution was
investigated. As illustrated in Fig. 7A (curve a), the pH at the anode
compartment of the black liquor decreases to 4.6 within 220 min
as compared to 35 min for the lignin solution, consistent with the
results observed in Fig. 3. With increased reaction duration, the
slope of the pH change in the cathode compartment (curve a of
Fig. 7B) gradually decreases due to a limitation in mass transport.
In order to improve the performance of the cell, the Na+ concentration at the cathode compartment was decreased and the current
density was elevated to 150 mA/cm2 , which greatly improved the
rate of the pH change (curves b of Fig. 7A and B). Interestingly, the
pH change performance of a 0.05 M NaOH cathodic solution system
(curves c of Fig. 7A and B) was increased only slightly, as compared
to the corresponding 0.05 M NaOH + 0.04 M Na2 SO4 solution systems (curves b of Fig. 7A and B). Consequently, although the mass
transport could be further improved, as illustrated in Fig. 6B, the cell
performance was also dependent on the conductivity of the system.

40
30
20
10
0

50

100

150

200

250

t / min
Fig. 9. Caustic recovery yields during electrochemical treatment of 10% black liquor
solution (anode compartment), cathode compartment: (a) 0.05 M NaOH + 0.20 M
Na2 SO4 at 100 mA/cm2 ; (b) 0.05 M NaOH + 0.04 M Na2 SO4 at 150 mA/cm2 ; (c) 0.05 M
NaOH at 150 mA/cm2 .

and led to further COD removal [35] in contrast to an acidic treatment. The vanillin and vanillic acid were detected using HPLC (data
not shown), in agreement with previous electrochemical studies
of lignin oxidation [28,35]. It should be noted that although the
current density for the 220 min-treatment was lower than other
electrochemical treatments, its reaction time was much longer,
leading to the generation of more intermediates and increased COD
removal.
Clearly, this signicant reduction in the COD value is due to the
oxidation and precipitation of lignin. Another signicant improvement inherent to the electrochemical treatment as compared to
traditional acidication methods is simultaneous caustic recovery
(Fig. 9 and Table 1). Owing to the electrolysis of water, there is
increasing OH that are generated at the cathode compartment,
facilitating the transport of Na+ from the black liquor solution

Fig. 8. The COD removal (A) values during the electrochemical treatment of 10% black liquor solution (anode compartment), cathodic compartment: (a) 0.05 M NaOH + 0.20 M
Na2 SO4 at 100 mA/cm2 ; (b) 0.05 M NaOH + 0.04 M Na2 SO4 at 150 mA/cm2 ; (c) 0.05 M NaOH at 150 mA/cm2 ; and the photographs (B) taken prior to and following acid
precipitation, and electrochemical treatment (0.05 M NaOH solution in the cathode at 150 mA/cm2 ).

W. Jin et al. / Electrochimica Acta 107 (2013) 611618

as a promising and cost-effective technique for the extraction of


lignin and recovery of caustic from black liquor.

Conductivity / Scm

-1

0.10
0.08

Acknowledgements
This work was supported by a Strategic Grant from the Natural Sciences and Engineering Research Council of Canada (NSERC).
A. Chen Acknowledges NSERC and the Canada Foundation of Innovation (CFI) for the Canada Research Chair Award in Material and
Environmental Chemistry.

0.06
0.04
0.02
0.00

617

References

Cycle numbers
Fig. 10. The conductivity change of Naon membrane after consecutive electrochemical treatment of the 10% black liquor solution (anode compartment), cathode
compartment 0.05 M NaOH at 150 mA/cm2 .

through the cation ion exchange membrane. There exist 50% Na+
in the black liquor solution that may be recovered in the electrochemical system developed in this study, contingent on mass
transport and electrochemical operational parameters.
Membrane durability and permeability properties are of great
importance for a membrane separation process [36]. Consequently,
the durability of the Naon membrane was investigated by monitoring its proton conductivity after the consecutive electrochemical
treatment of the black liquor solution. As shown in Fig. 10, the
initial membrane conductivity was 0.073 S cm1 , which is consistent with the value reported in the literature [37]. The membrane
conductivity was slightly decreased after ve cycles of the consecutive electrochemical treatment of the black liquor solution. This is
likely due to the adsorption of lignin and lignin derivatives on the
membrane surface.
4. Conclusions
In this study, we have developed a novel membrane-assisted
electrochemical approach for the precipitation of lignin from black
liquor. Without the addition of acid or CO2 , the pH in the black
liquor solution was lowered to 4.7 due to water electrolysis, leading to pH-dependent lignin precipitation. Simultaneously, Na+ ion
traverses the membrane to the cathode compartment to balance
the OH that is produced during the cathodic reaction, facilitated
caustic recovery from the black liquor solution. Owing to lignin
precipitation and oxidation, more than 70% of the COD value in
the black liquor solution was decreased at the set pH, which indicated a marked improvement of COD reduction compared with the
conventional acidic precipitation. Our study has also shown that
the electrochemical cell performance is signicantly inuenced by
the condition of the electrolytes. Due to the lignin-derived intermediate adsorption at the electrode surface, the electrochemical
oxygen evolution was partially suppressed, leading to a sluggish
pH transition in the system. With decreasing the cation concentration at the cathode compartment, mass transport was improved,
and thus Na+ transport through the membrane to the cathode compartment was enhanced. Otherwise, the H+ generated at the anode
compartment would competitively transport to the cathode compartment to frustrate the cell performance. Moreover, it has shown
that the solution conductivity and the applied current density also
play important roles in altering the pH of the system. In summary,
this novel membrane-assisted electrochemical approach may serve

[1] A. Demirbas, Recovery of oily products from organic fraction of black liquor via
pyrolysis, Energy Sources 30 (2008) 1849.
[2] M. Cardoso, E.D. Oliverira, M.L. Passos, Chemical composition and physical
properties of black liquors and their effects on liquor recovery operation in
Brazilian pulp mills, Fuel 88 (2009) 756.
[3] X.Y. Zhao, J.P. Cao, K. Morishita, J. Ozaki, T. Takarada, Electric double-layer
capacitors from activated carbon derived from black liquor, Energy Fuels 24
(2010) 1889.
[4] A. Tejado, C. Pena, J. Labidi, J.M. Echeverria, I. Mondragon, Physicochemical characterization of lignins from different sources for use in
phenolformaldehyde resin synthesis, Bioresource Technology 98 (2007)
1655.
[5] F.G. Calro-Flores, J.A. Dobado, Lignin as renewable raw material, ChemSusChem
3 (2010) 1227.
[6] D.H. Lataye, I.M. Mishra, I.D. Mall, Removal of pyridine from aqueous solution
by adsorption on bagasse y ash, Industrial & Engineering Chemistry Research
45 (2006) 2830.
[7] A.L. Kohl, Black liquor gasication, Canadian Journal of Chemical Engineering
64 (1986) 299.
[8] G. Thompson, J. Swain, M. Kay, C.F. Forster, The treatment of pulp and paper
mill efuent: a review, Bioresource Technology 7 (2001) 275.
[9] C. Yang, Y. Niu, H. Su, Z. Wang, F. Tao, X. Wang, H. Tang, C. Ma, P. Xu, A
novel microbial habitat of alkaline black liquor with very high pollution load:
microbial diversity and the key members in application potentials, Bioresource
Technology 101 (2010) 1737.
[10] L.V. Da-Re, L. Papinutti, Black liquor decolorization by selected white-rot fungi,
Applied Biochemistry and Biotechnology 165 (2011) 406.
[11] D.R. Dimmel, J.J. Bozell, Pulping catalysts from lignin, TAPPI Journal 74 (1991)
239.
[12] A. Garg, I.M. Mishra, S. Chand, Effectiveness of coagulation and acid precipitation processes for the pre-treatment of diluted black liquor, Journal of
Hazardous Materials 180 (2010) 158.
[13] M. Nagy, M. Kosa, H. Theliander, A.J. Ragauskas, Characterization of CO2 precipitation Kraft lignin to promote its utilization, Green Chemistry 12 (2010)
31.
[14] M. Kosa, H. Ben, H. Theliander, A.J. Ragauskas, Pyrolysis oils from CO2 precipitated Kraft lignin, Green Chemistry 13 (2011) 3196.
[15] Z. Wei, Y. Yang, R. Yang, C. Wang, Alkaline lignin extracted from furfural residues
for pH-responsive Pickering emulsions and their recyclable polymerization,
Green Chemistry 14 (2012) 3230.
[16] F. Ohman, H. Theliander, Filtration properties of lignin precipitated from black
liquor, TAPPI Journal 6 (2007) 3.
[17] H. Lout, B. Blackwell, V. Uloth, Lignin recovery from kraft black liquor: preliminary process design, TAPPI Journal 74 (1991) 203.
[18] H.R. Ghatak, Electrolysis of black liquor for hydrogen production: some initial
ndings, International Journal of Hydrogen Energy 31 (2006) 934.
[19] M. Zaied, N. Bellakhal, Electrocoagulation treatment of black liquor from paper
industry, Journal of Hazardous Materials 163 (2009) 995.
[20] H.R. Ghatak, Reduction of organic pollutants with recovery of value-added
products from soda black liquor of agricultural residues by electrolysis, TAPPI
Journal 7 (2009) 4.
[21] M.A. Blanco, C. Negro, J. Tijero, A.C.M.P. De Jong, D. Schmal, Electrochemical
treatment of black liquor from straw pulping, Separation Science and Technology 31 (1996) 2705.
[22] K. Chanworrawoot, M. Hunsom, Treatment of wastewater from pulp and paper
mill industry by electrochemical methods in membrane reactor, Journal of
Environmental Management 113 (2012) 399.
[23] K.A. Mauritz, R.B. Moore, State of understanding of Naon, Chemical Reviews
104 (2004) 4535.
[24] R. Tolba, M. Tian, J. Wen, Z.-H. Jiang, A. Chen, Electrochemical oxidation of
lignin at IrO2 -based oxide electrodes, Journal of Electroanalytical Chemistry
649 (2010) 9.
[25] L.A. Da Silva, V.A. Alves, M.A.P. Da Silva, S. Trasatti, J.F.C. Boodts, Oxygen evolution in acid solution on IrO2 + TiO2 ceramic lms. A study by impedance,
voltammetry and SEM, Electrochimica Acta 42 (1997) 271.
[26] L.K. Xu, J.D. Scantlebury, Microstructure and electrochemical properties of
IrO2 Ta2 O5 -coated titanium anodes, Journal of the Electrochemical Society 150
(2003) B254.

618

W. Jin et al. / Electrochimica Acta 107 (2013) 611618

[27] L. Xu, Y. Xin, J. Wang, A comparative study on IrO2 Ta2 O5 coated titanium
electrodes prepared with different methods, Electrochimica Acta 54 (2009)
1820.
[28] M. Tian, J. Wen, D. MacDonald, R.M. Asmussen, A. Chen, A novel approach
for lignin modication and degradation, Electrochemistry Communications 12
(2010) 527.
[29] W. Jin, H. Du, S. Zheng, H. Xu, Y. Zhang, Comparison of the oxygen reduction
reaction between NaOH and KOH solutions on a Pt electrode: the electrolytedependent effect, Journal of Physical Chemistry B 114 (2010) 6542.
[30] W. Jin, M.S. Moats, S. Zheng, H. Du, Y. Zhang, J.D. Miller, Indirect electrochemical Cr(III) oxidation in KOH solutions at an Au electrode: the role of oxygen
reduction reaction, Journal of Physical Chemistry B 116 (2012) 7531.
[31] J. Rossmeisl, A. Logadottir, J.K. Nrskov, Electrolysis of water on (oxidized) metal
surfaces, Chemical Physics 319 (2005) 178.
[32] R.A. Rozendal, H.V.M. Hamelers, C.J.N. Buisman, Effects of membrane cation
transport on pH and microbial fuel cell performance, Environmental Science
and Technology 40 (2006) 5206.

[33] R.F.D. Costa, J.Z. Ferreira, C. Deslouis, Electrochemical study of the interactions
between trivalent chromium ions and Naon peruorosulfonated membranes, Journal of Membrane Science 215 (2003) 115.
[34] P.K. Sow, S. Sant, A. Shukla, EIS studies on electro-electrodialysis cell for concentration of hydriodic acid, International Journal of Hydrogen Energy 35 (2010)
8868.
[35] K. Pan, M. Tian, Z.H. Jiang, B. Kjartanson, A. Chen, Electrochemical oxidation of
lignin at lead dioxide nanoparticles photoelectrodeposited on TiO2 nanotube
arrays, Electrochimica Acta 60 (2012) 147.
[36] L. Alvarado, I.R. Torres, A. Chen, Integration of ion exchange and electrodeionization as a new approach for the continuous treatment of hexavalent
chromium wastewater, Separation and Purication Technology 105 (2013)
55.
[37] Y. Sone, P. Ekdunge, D. Simonsson, Proton conductivity of Naon 117 as measured by a four-electrode ac impedance method, Journal of the Electrochemical
Society 143 (1996) 1254.

Potrebbero piacerti anche