Sie sind auf Seite 1von 42

Accepted Manuscript

The electrochemical degradation of ciprofloxacin using a SnO2-Sb/Ti anode:


Influencing factors, reaction pathways and energy demand

Ying Wang, Chanchan Shen, Manman Zhang, Bo-Tao Zhang, Yan-Ge Yu

PII: S1385-8947(16)30346-1
DOI: http://dx.doi.org/10.1016/j.cej.2016.03.093
Reference: CEJ 14945

To appear in: Chemical Engineering Journal

Received Date: 27 October 2015


Revised Date: 15 March 2016
Accepted Date: 19 March 2016

Please cite this article as: Y. Wang, C. Shen, M. Zhang, B-T. Zhang, Y-G. Yu, The electrochemical degradation of
ciprofloxacin using a SnO2-Sb/Ti anode: Influencing factors, reaction pathways and energy demand, Chemical
Engineering Journal (2016), doi: http://dx.doi.org/10.1016/j.cej.2016.03.093

This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers
we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and
review of the resulting proof before it is published in its final form. Please note that during the production process
errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.
The electrochemical degradation of ciprofloxacin using a SnO2-Sb/Ti anode:

Influencing factors, reaction pathways and energy demand

Ying Wang1 *, Chanchan Shen1, Manman Zhang1, Bo-Tao Zhang2, Yan-GeYu1

1
The Key Laboratory of Water and Sediment Sciences, Ministry of Education, School

of Environment, Beijing Normal University, Beijing 100875, P.R. China

2
College of Water Sciences, Beijing Normal University, Beijing 100875, P.R. China

Abstract

The electrochemical oxidation of ciprofloxacin (CIP) using a SnO2-Sb/Ti

electrode was systematically investigated. The effects of the current density, initial

concentration of CIP, and initial pH were evaluated. The results showed that

electrochemical oxidation, using a SnO2-Sb/Ti electrode was highly effective for the

degradation of CIP. After 120 min, the removals of CIP (50 mg/L), COD and TOC at

a current density of 30 mA/cm2 were about 99.5%, 86.0%, and 70.0%, respectively.

The reaction followed a first-order kinetics model. The current density and initial

concentration of the CIP exerted a prominent effect on the degradation of CIP, COD,

and electrical energy demand, while the initial pH had no effect. The electrochemical

degradation pathways, of CIP in aqueous solution, were studied using ion

Corresponding author
*

Tel.: +86-10-5880 2851


E-mail: yingwang@bnu.edu.cn
1
chromatography and liquid chromatography coupled with mass spectrometry

(LC-MS). Three major degradation pathways were proposed: oxidation of the

piperazine ring, hydroxylation of the quinolone moiety, and defluorination (OH/F

substitution). Inorganic N compounds were and . F was reduced to F-.

Keywords: Electrochemical degradation; Antibiotics; SnO2-Sb/Ti anode;

Intermediates; Energy demand

1. Introduction

In recent years, an increasing amount of antibiotics have been found in various

aquatic environments including surface water, ground water, and sewage treatment

plant effluents, which has become an emerging environmental concern [1]. Most of

these antibiotics cannot be removed effectively by wastewater treatment plants

(WWTPs) due to their poor biodegradability, which consequently results in an

accumulation of antibiotics in water. Antibiotics, even in trace amounts, have the

potential to increase the antibiotic resistance of pathogens within humans, animals,

and threatens the functionality of an ecosystem [2]. Fluoroquinolones (FQ) are a

family of synthetic, broad-spectrum antibacterial compounds that are used to combat

both human and veterinary diseases [3]. Over the last few decades, due to their annual

global sales and therapeutic versatility, they are among the five classes of antibiotics

(-lactam, macrolides, fluoroquinolones, sulfonamides, and tetracyclines) that are

2
frequently detected in wastewater and surface waters in relatively high concentrations

[4]. Thus, determining how to eliminate them effectively from waste and drinking

water is an urgent priority.

Advanced oxidation processes (AOPs), which are characterized by the formation

of a hydroxyl radical, which is a non-selective oxidant, have been widely applied to

the treatment of biorefractory compounds [5]. Within these processes, the

electrochemical advanced oxidation processes (EAOPs) have been paid more

attention. Sirs et al. outlined the EAOPs technologies and reviewed some

remediation of water pollution caused by pharmaceutical residues based on EAOPs [6,

7]. EAOPs possess high oxidation performance, ease of control, environmental

friendliness, high energy efficiency, and cost effectiveness [8]. To date, EAOPs have

been used to degrade various antibiotics such as semi-synthetic beta-lactam antibiotics,

sulfonamides, tetracyclines, and so on [9-11]. However, relatively few studies exist

regarding the treatment of quinolones using electrochemical methods. Several studies

reported that the degradation of enrofloxacin and ofloxacin on a boron-doped

diamond (BDD) electrode and the oxidation of ciprofloxacin, flumequine, and

enrofloxacin by the electro-Fenton method were effective [12-16]. However, the BDD

electrode is expensive and the electro-Fenton process requires acidic conditions [17,

18]; thus, their large-scale application is limited. Recently, Sb-doped SnO2 electrodes

have been widely applied to treat different organic pollutants containing antibiotics.

Although the life span of Sb-doped SnO2 electrodes is relatively shorter when

3
compared with other electrodes, numerous advantages exist, such as high oxygen

evolution over-potential, excellent electro-catalytic performance, cost efficiency and

ease of preparation [19-23]. These properties increased its application to disposal of

organic pollutants. Hence, it is possible to degrade fluoroquinolones effectively using

SnO2-Sb/Ti electrode to achieve the dual goals of low cost and mild conditions.

CIP, a second generation quinolone, is the most widely prescribed quinolone in

Europe [3]. It has been repeatedly found in relatively high concentrations ranging

from g/L to ng/L levels in hospital effluents, in secondary wastewater effluents, and

in raw drinking water. Surprisingly, CIP was found in waste effluents from

pharmaceutical manufacturers at concentrations of up to 31 mg/L [24]. Due to its high

concentration in water and harm to both humans and the ecosystem, it was chosen as a

model pollutant in this study.

The principal goal of this study was to provide an evaluation of the effect of

degrading CIP using SnO2-Sb/Ti electrode. The main influencing factors, including

the applied current density, the initial CIP concentration, and the pH, were discussed

systematically. To comprehensively understand the mechanism of CIP degradation,

the intermediates, including both released inorganic ions and organics, were identified.

Finally, the electrical energy consumption per order of magnitude (EE/O) of CIP

degradation by SnO2-Sb/Ti electrode was assessed.

2. Materials and methods

2.1 Chemicals

4
Ciprofloxacin (CIP>98%) was purchased from Tokyo Chemical Industry

development Co., Ltd. (Shanghai, China). All of the chemicals employed were of

analytical grade and were used without further purification. All aqueous solutions

were prepared with ultrapure water.

2.2 Preparation of SnO2-Sb/Ti electrode

Titanium mesh (50mm 30 mm1 mm) was used for the substrates. The titanium

mesh has 63 diamond holes. These diamond holes were distributed in 21 rows with 3

dimaond holes in each row. They were thoroughly polished with a series of different

abrasive papers to a mirror-like finish. They were degreased in a 5% NaOH solution,

which was boiled 1 h, and then etched in boiling a 10% oxalic acid solution for 2 h.

Finally, they were washed with ultrapure water. The SnO2-Sb/Ti electrode was

prepared by the sol-gel method [25]. First, citric acid (CA) was dissolved in ethylene

glycol (EG) at 60 C for 30 min. Next, the solution was raised to a temperature of

90 C. Then, SnCl44H2O and SbCl3 were added to the solution. The solution was

maintained at 90 C for 30 min. The final molar ratio of the sol-gel was 140: 30: 9: 1

(EG: CA: SnCl44H2O: SbCl3). The pretreated titanium meshes were dipped into the

sols. Then, the coating was dried at 140 C for 10 min, followed by annealing at

500 C in a muffle oven for 10 min. This procedure was repeated 15 times, and the

last coating was annealed at 500 C for 1 h.

2.3 Electrochemical experiments

The electrochemical experiments were conducted in a 250 mL beaker stirred

5
with a magnetic stirrer. The temperature was maintained at 30 C. The SnO2-Sb/Ti

electrode (30 mm 50 mm1 mm) and Ti plate (30 mm 50 mm1 mm) served as the

anode and cathode, respectively. The distance between the anode and cathode was 2

cm. At precise time intervals, fixed amounts of the reaction solution were withdrawn

and analyzed by HPLC. At the electrolysis time of 30 min, sample was withdrawn to

analyze intermediates by LC-MS.

The removals of CIP, COD and TOC were calculated according to the following

equation:

-
(1)

Where C0 and Ct are the initial concentration and concentration at time t,

respectively.

Kinetic parameters analysis

In electrochemical oxidation, the reaction between CIP or COD with OH is

expressed as follows:

or (2)

(3)

Then the oxidation rate of CIP or COD can be written as:

With as the absolute rate constant of the second order reaction.

The operating parameters were constant throughout the experiments, as a result

6
the production of hydroxyl radicals were constant as well. The second order reaction

can be approximated to a pseudo first order. Considering these conditions, the general

rate equation can be simplified to:

With

By plotting Ln ([CIP]0/[CIP]t) or Ln ([COD]0/[COD]t) as a function of reaction

time t, through linear regression, the kapp can be derived from the slopes of the straight

lines.

The actual service lifetime of electrodes is relatively long due to application at

low current density, which is difficult to evaluate. Correa-Lozano et al. [26] have

proposed an empirical relationship between the actual service life (t1) and the

accelerated service life (t2). The equation can be used to estimate the actual lifetime of

the electrode in different current densities.

( (8)

Where, i1 and i2 are actual current density and accelerated current density,

respectively.

The instantaneous current efficiency (ICE) was calculated by [27]:

ICE = FV (9)

Where CODt and CODt+ t are the COD at time t and t+t (g O2/dm3),

respectively; F is the Faraday constant (96487 C/mol); V is the volume of the

7
electrolyte (dm3) and I is the current (A).

The average current efficiency (ACE) was then calculated by [27]:

ACE = (10)

Where E he n an ane en eff en y an he a n f he

electrochemical treatment (120 min).

An electrical energy consumption per order of magnitude (EE/O) was

recommended to evaluate the energy consumption of the process. EE/O is defined as

the number of watt-hours, of electrical energy, required for degrading a pollutant and

reducing its concentration by one order of magnitude (90%) in 1 L of contaminated

water [28]. The EE/O was calculated by the following equation:

EE/O (Wh/L) = (11)

Where Ucell is average cell voltage (V), I is the current (A), t is the time needed

to degrade the CIP, V is the volume (L), C0 is the initial concentration (mg/L), and Ct

is the final concentration (mg/L).

In addition to EE/O, one of the practical yardsticks in an industrial environment

is the energy efficiency, which gives a description of the efficiency of a given process

to determine its economic viability. The energy efficiency of degradation is defined as

the ppm of organic contamination degraded in a given volume of solution per Wh of

electrical energy [29]. The energy efficiency was calculated by the following

equation:

Energy efficiency (mg/Wh) = (12)


8
Where m is the amount of contaminant degraded (mg), U is the average

electrolysis voltage (V), I is the electrolysis current (A), t is the time for degradation

of the amount of contaminant (h).

2.4 Electrode characterization

The surface morphology and crystal structure of the electrodes were

characterized using scanning electron microscopy (SEM, S4800, Hitachi, Japan) with

an accelerating voltage of 15 kV and X-ray diffraction (XRD; Thermo ARL

S NTAG XTRA, Ne he lan ng K a a n( kV, mA , e pe vely.

The electrochemical properties were determined by a CHI 660 (Shanghai

Chenhua Instrument Co., China) electrochemical workstation. The SnO2-Sb/Ti

electrode (10 mm 20 mm), a platinum plate (10 mm 20 mm), and an Ag/AgCl

electrode were employed as the working, counter, and reference electrodes,

respectively.

Accelerated life test [30] was carried out under electrolysis at constant current

density of 100 mA/cm2 in 0.5 mol/L H2SO4 solution. The anode potential was

measured as a function of time, and it was considered that the electrode is deactivated

when the potential increases 5 V from its initial value.

2.5 Analytical instruments

The concentrations of CIP were measured on a high-performance liquid

chromatograph (HPLC-UV, Dionex U3000, USA) equipped with a WpHC-18 column

(150 mm4.0 mm . ., m). The mobile phase contains 75% water (0.1% phosphoric

9
acid) and 25% methanol. The flow rate was kept at 1.0 mL/min, and the injection

volume was 20 L. The UV absorption of CIP was set at 274 nm.

An ion chromatograph system (DX-600, Dionex, USA) was used to determine

the inorganic ions.

A multi N/C UV analyzer (Analytic Jena, Germany) was used to analyze the total

organic carbon (TOC) and total nitrogen (TN).

COD was determined according to the standard potassium dichromate digestion

method (GB11914-89) of the Ministry of Environmental Protection of China [31].

The concentration of persulfate was measured by a spectrophotometric method

using potassium iodide [32].

The organic intermediates were identified by a system of liquid chromatography

coupled with a mass spectrometer (LC-MS) (Waters ACQUITY UPLC/Xevo G2 Q

TOF) equipped with an ACQUITY UPLC BEH C18 column (50 mm2.1 mm i.d.,1.7

m . The m b le pha e wa mp e f wa e (A an a e n le (B a a fl w a e

of 0.3 mL/min. The mobile phase progressed from constant 10% B (initial conditions)

for 1 min to a linear gradient of 100% B over 10 min. For MS analysis, ionization was

conducted through positive-mode electrospray ionization (ESI+) with a full scan from

m/z 50 to m/z 1000 under the following conditions: capillary potential: 3.5 kV; cone

voltage: 30 V; source temperature: 120 C; desolvation temperature: 300 C.

3. Results and discussion

3.1 Characteristics of the Ti/SnO2-Sb electrode

10
3.1.1 SEM of Ti/SnO2-Sb electrode

Fig. 1

The SnO2-Sb/Ti electrode was studied by a scanning electron microscope to

investigate its surface morphology. The composition of the precursor solution and the

preparation conditions play key roles in the surface morphology of the electrode

coating. Fig. 1 (a) illustrates that the film is compact, with rare cracks or defects

observed. Fig. 1 (b) shows that the coating particles of the SnO2-Sb/Ti electrode are

uniform in size, with diameters ranging from 20 nm to 100 nm. This nanostructure of

the electrode has a larger surface area, which consequently may provide more active

sites for electrochemical oxidation.

3.1.2 XRD of SnO2-Sb/Ti electrode

XRD was used to study the composition structures of the SnO2-Sb/Ti electrode.

Fig. 1 (c) presents the XRD patterns of the electrode surface, which only demonstrates

a series of diffraction peaks of the SnO2 structure. The peak positions agree well with

the reflections of cassiterite indicating a rutile-type structure, and its characteristic

peak l a e a . , 33.9, 3 .9, .8, .8, .9 an 8. we e a gne he

lattice planes (110), (101), (200), (211), (220), (310) and (321) of SnO2, respectively.

The m n m m y all e ze wa al la e be nm a ng S he e

formula as follows:

D = R /( (12)

11
n he f m la, he ame e f he pa le , R S he e n an (0.89),

he n en waveleng h ( . nm , ( a he peak w h a half he gh an

( he ff a n angle.

From the XRD spectra, no clear diffraction peaks corresponding to the oxides of

antimony were found because of either the low doping level of Sb or the incorporation

of the doping Sb into SnO2. In addition, no obvious diffraction peaks corresponding to

titanium dioxide were detected, indicating that the titanium substrate was well

covered by the tin dioxide coating.

3.2. Electrochemical degradation of CIP

3.2.1 Performance of the system

Fig. 2

The electrochemical characteristic of the SnO2-Sb/Ti electrode was investigated

using cyclic voltammetry (CV). The CV curves of the SnO2-Sb/Ti electrode were

performed between 0 and 2 V (SCE) in the absence of CIP and in the presence of 50

mg/L CIP. In Fig. 2(a), when 50 mg/L of CIP was added to the supporting electrolyte,

no additional peaks were found in the CV curve. This phenomenon may indicate that

the oxidation of CIP did not occur on the surface of the electrode or the oxidation

could occur at the zone of oxygen evolution, because the current densities used in this

work are higher.

Fig. 2(b) shows the CIP (50 mg/L), COD (89.2 mg/L), and TOC (30.9 mg/L)

12
removals of the system. It could be seen that after 120 min, the removals of CIP, COD

and TOC were 99.5%, 86.0% and 70.0%, respectively. The result showed that the

SnO2-Sb/Ti electrode was efficient in electrochemical oxidative degradation of CIP.

He et al. [33] and Wu et al. [34] studied the photocatalytic degradation of CIP by

yttrium-doped BiOBr (initial concentration: 10 mg/L, volume: 100 mL) and

heteropolyacid / TiO2 / fly-ash-cenosphere (initial concentration: 20 mg/L, volume: 50

mL), respectively. They found that the degradation reached 78% after 6 h of

irradiation by yttrium-doped BiOBr and 60% after 50 min by heteropolyacid / TiO2 /

fly-ash-cenosphere, respectively. In comparison, the SnO2-Sb/Ti electrode possessed a

higher electrochemical oxidation capability for CIP (initial concentration: 50 mg/L,

volume: 250 mL). Chen et al. [13] used BDD electrodes to degrade ofloxacin.

Ofloxacin was nearly fully degraded at applied current density of 20 mA/cm2 after

150 min of electrolysis time (initial concentration: 50 mg/L, volume: 200 mL, BDD

anode surface area; 1 cm2). The results indicated that BDD electrode showed better

fluoroquinolones removal efficiency than SnO2-Sb/Ti electrode. However, BDD

electrode is so expensive that its application is limited. Yahya et al. [35] reported

electro-fenton removal of CIP with a Pt anode and a carbon felt cathode. 0.15 mmol/L

CIP with a volume of 230 mL was completely removed after 10 min, while about 19%

of TOC was degraded after 120 min. Compared with electro-fenton, anodic oxidation

using SnO2-Sb/Ti electrode had a lower CIP removal efficiency but a higher

mineralization of CIP. Electro-fenton requires acidic condition. Pt and BDD

electrodes were more expensive. These would increase operative difficulty and cost.

13
As a result, SnO2-Sb/Ti electrode was more suitable for decomposition of CIP in

terms of degradation efficiency and application cost. The electrochemical stability is

an important factor for SnO2-Sb/Ti electrode. Fig. 2(c) shows the accelerated service

life test result of the SnO2-Sb/Ti electrode at current density of 100 mA/cm2 in 0.5 M

H2SO4 solution at 30 C. It is observed that the potential of the electrode was stable at

the beginning period before 60 h. Between 60 and140 h, the potential of the electrode

increased slowly to about 4.0 V. After 140 h, the SnO2-Sb/Ti electrode undergone a

sharp rise to 8.0 V at 160 h, indicating the deactivation of the SnO2-Sb/Ti electrode.

Therefore, the accelerated lifetime for the SnO2-Sb/Ti electrode was 160 h.

According to equation (8), the actual service lifetime of the SnO2-Sb/Ti electrode

was 640 h under the current density of 50 mA/cm2. Compared with the study of

Zhang et al. [30], the actual service lifetime of SnO2-Sb/Ti electrode in our study was

longer than the SnO2-Sb/Ti electrode (380 h), but lower than the carbon nanotube

modified SnO2-Sb/Ti electrode (1816 h) in their research.

3.2.2 Effect of the applied current density

Fig. 3

It is well known that the applied current density plays an important role in the

electrochemical degradation of organic compounds because the amount of hydroxyl

radical is related to the applied current density. Fig. 3(a) shows the effect of the

applied current density on the degradation of 50 mg/L CIP. Clearly, the degradation

14
was significantly affected by the current density. With an increase in the applied

en en y ( 30 mA/cm2), the removal was enhanced. However, when the

applied current density continued to increase to 40 mA/cm2, the removals of CIP and

COD decreased. This was due to a higher applied current density that not only

increased oxygen evolution (Eq. (13)), but also enhanced the decomposition of the

electrolyte (Eq. (14). After 120 min electrolysis time, concentration reached

.9, .3, 3. an . mol/L at the applied current density of 10, 20, 30 and 40

mA/cm2, respectively. These two reactions would compete with the degradation of the

contaminant [36, 37].

e- (13)

- -
S S 8 e- (14)

As shown in Fig. 3(a), a plot of Ln (C0/C) versus time (t) presented a straight line.

The apparent rate constant k and correlation coefficient R2 are summarized in Tab 1.

According to the regression analysis (Tab. 1), all of the correlation coefficients were

higher than 0.95, implying the electrochemical degradation of CIP followed

pseudo-first-order kinetics. This result was in agreement with other studies [34]. The

k values for the degradation of CIP and COD at a current density of 30 mA/cm 2 were

approximately 2 times that at a current density of 10 mA/cm2.

As shown in Tab 1, the ACE decreased as the current density increased at the

corresponding operating times, which may be due to competitive electrode reactions,

such as enhanced oxygen evolution. It revealed that a low current density is beneficial

for achieving high current efficiency [38].

15
3.2.3 Effect of the initial CIP concentration

The effect of the initial CIP concentration on its electrochemical degradation was

studied. In Fig. 3(b), as a whole, the electrochemical degradation rate of CIP and

COD increased with a decrease in the initial CIP concentration. The result was similar

with the degradation of ofloxacin on a BDD electrode [13] and the degradation of CIP

by photolysis [39]. This phenomena was most likely attributed to the fact that more

intermediates were produced as the reaction proceeded at higher CIP concentrations,

which may compete with the CIP itself for the active sites. This can be demonstrated

by the change in the UV absorption spectra (Fig. 3(c)). It can be seen that the

intensities of the absorbance peaks at 274 and 330 nm dropped continuously and

disappeared completely after 120 min at an initial concentration of 30 mg/L. This was

a clear sign that the CIP concentration was progressively decreasing and that the CIP

was completely degraded. This result was in accordance with that reported by Liu et

al. and Wu et al. [34, 40]. A similar trend was also obtained for the initial CIP

concentrations at 10 and 50 mg/L. When the initial CIP concentration was 100 mg/L,

there was an increasing peak at 400 nm, which was not observed for the initial CIP

concentrations of 10, 30, and 50 mg/L. This result indicated that some intermediates

accumulated at the higher initial CIP concentrations, which then competed with the

CIP to decrease the CIP removal efficiency. The increasing UV absorption peak at 400

nm was also found in the degradation of CIP by ZnO nanoparticles under ultraviolet

radiation in the study by Kemary et al. [41].

Although the removal efficiency of CIP decreased with the increasing initial CIP

16
concentration, absolute CIP removal amount increased. As the CIP concentration

increased from 10 to 100 mg/L, the absolute CIP removal amounts increased from

about 10.0 mg/L to 69.2 mg/L after 60 min electrolysis. This demonstrated that

SnO2-Sb/Ti electrode remained to be high efficiency at high concentration of CIP.

This phenomenon was consistent with that of the oxidation of ofloxacin on a BDD

electrode by Chen et al. [13]. A possible reason included that the increase of initial

CIP concentration enhanced its concentration gradient and mass transfer across the

diffusion layer, which thus increased its degradation on electrode. As a consequence,

the total amount of degraded CIP was greater with the increasing of the initial CIP

concentration.

The value of ACE increased with an increase in the initial CIP concentrations.

Some studies found that the higher initial organic concentrations resulted in an

increase of A E. A h ghe n al n en a n , he a al have a g ea e

chance to react rapidly with the organic matter, which led to the reduction of the

oxygen-evolution side reaction [27].

3.2.4 Effect of the pH

The effect of pH on the CIP and COD degradation was measured. The results are

shown in Fig. 3(d) and Tab. 1. No significant difference in the k values and ACE were

observed with an increase of pH from 3 to 11. The results indicated that there was a

slight effect of pH on the CIP degradation rate. This suggested that the

electrochemical degradation of CIP by a SnO2-Sb/Ti electrode could be applied in a

wide pH range. A similar result was observed on the oxidation of Acid Red 73 by a

17
SnO2-Sb/Ti electrode [42].

Specifically, the degradation rates of both CIP and COD at a lower initial pH

were slightly higher than that at a higher pH. At an initial pH of 3, the k value and

ACE were maximal. The effect of the initial pH on the degradation of CIP was likely

the result of two effects: the oxygen-evolution side reaction and the speciation of CIP.

Lower pH could increase the oxygen over-potential [43]. Therefore, the

oxygen-evolution side reaction (Eq. (13)) decreased and more hydroxyl radical was

generated at lower pH, which enhanced the removal of CIP. Additionally, CIP is

zwitterionic compound showing two pKa values (pKa1 = 6.1; pKa2 = 8.7). So, CIP

could exist as different species: cationic (pH < 6.1), zwitterionic (6.1< pH < 8.7) or

anionic (pH > 8.7) form at different pH solution. Under acidic conditions, CIP is

positively charged, which makes it much difficult to be adsorbed on the electrode

surface. This would decrease the electromigration mass transfer. The effect would

reverse under alkaline condition. In acidic medium, the inhibition effect of

oxygen-evolution side reaction was stronger than the decrease of electromigration

mass transfer, resulting in the quick degradation at lower pH. Under strong alkaline

conditions, electrostatic attraction was stronger, which resulted in the relatively faster

degradation at pH of 11 than that at pH of 9.

Tab. 2

3.2.5 The evolution of the electrochemical oxidation intermediates and

degradation pathways of CIP


18
3.2.5.1 Evolution of inorganic intermediates

Fig. 4

The complete mineralization of CIP leads to the conversion of a fluorine

substituent and three atoms of nitrogen into inorganic ions. These species could

provide important information about the oxidation mechanisms. In Fig. 4, the results

showed that F-, , and were formed as the main inorganic ion products.

Similar main inorganic ion products were also reported by others [14, 15, 35]. The

inorganic ions generated indicated that the possible cleavage of C-F and C-N bonds in

the CIP molecule after being attacked by OH. The amount of F- gradually increased

with electrolysis. After 120 min, the concentration of F- reached approximately 2.0

mg/L (71.9% of initial F). The deficient initial fluorine mass balance suggested the

undetected fluorine was in organic F derivatives. Correspondingly, organic F

derivative concentrations decreased with electrolysis time. During mineralization

-
process, the production of inorganic N ions were N -N and N 3 -N. No was

detected in any experiment solution. This may be because was extremely

unstable and was rapidly transformed into or during the electrochemical

process. The concentration of N -N increased steadily during electrolysis and

reached a final value of 2.5 mg/L (40.2% of initial N) after 120 min reaction. The

-
N 3 -N increased at first, reaching a maximum of 1.4 mg/L, and then decreased to 1.0

mg/L (16.1% of initial N), which indicated that nitrate was an intermediate product.

19
As reported, in the electrochemical process, can be reduced to or gas,

such as N2, NH3, N2O or NO2, at the cathode [15]. In addition, the total nitrogen (TN)

concentration of electrolyzed solutions were determined. TN concentration decreased

gradually to 4.6 mg/L (73.1% of initial TN) with the increasing electrolysis time. This

suggested that part of N can be lost as volatile organic compounds containing nitrogen

or reduced to gas such as N2, NH3 or N2O [14]. Finally, it should be noted that organic

N compound concentrations decreased with reaction time.

3.2.5.2 Intermediates and the degradation pathways of CIP

To elucidate the degradation mechanism of CIP in the electrochemical system

and thus to illustrate a possible reaction pathway, a mass spectroscopy study was

carried out to determine the intermediates in aqueous solution. The intermediates

during CIP degradation by the electrochemical system were identified by the peaks

formed using HPLC-MS. Six intermediate compounds were identified as shown in

Tab. 2. According to previous studies [35, 39, 44-46], in advanced oxidation, CIP was

most likely degraded at the quinolone moiety and the piperazine ring. Thus, three

major possible pathways for the electrochemical oxidation of CIP were proposed, as

shown in Fig. 5.

Pathway 1 was the oxidative degradation of the piperazinyl moiety. In our study,

the partial breakdown of the piperazine moiety led to the formation of m/z 291. The

decarbonylation of m/z 291 formed the aniline m/z 263, in which the piperazinyl

substituent of CIP was completely destroyed [45].

Pathway 2 was proposed to be the hydroxylation process. An et al. [47]

20
calculated the frontier electron densities (FEDs) of CIP to predict the reaction sites for

hydroxyl radical attack. According to their study, OH was most likely to attack on

quinolone ring due to calculation. m/z 348 was assigned to be the product of such a

hydroxylation, with OH attacking quinolone moiety. Then, the quinolone ring broke,

resulting in the formation of m/z 366.

Pathway 3 is the well-known degradation pathway of defluorination [46, 48].

During defluorination, the attack of OH at the carbon-fluorine site resulted in the

substitution of fluorine by a hydroxyl group and led to a germinal fluorohydrin with

the loss of HF. The product would be the compound with m/z 330. Then, after losing

two H and gaining one O atom from the piperazinyl group of compound m/z 330, the

keto-derivative m/z 344 formed.

Tab. 2

Fig. 5

3.2.6 The electrical energy consumption

Fig. 6

For the electrochemical degradation, energy demand is an important factor that

determines the economic feasibility of the process. The major operating cost is

21
associated with electrical energy consumption during the electrochemical degradation

process. Hence, it is necessary to compare the process efficiency. One such measure is

the electrical energy per log order (EE/O) of compound destruction. Mahdi-Ahmed

and Chiron studied the degradation of CIP by UV/persulfate (PDS),

UV/peroxymonosulfate (PMS) and UV/H2O2. The electrical energy per order of CIP

degraded by UV/H2O2, UV/PMS and UV/PDS was 22.6 Wh/L, 14.8 Wh/L, and 8.3

Wh/L, respectively, for an n al n en a n f mol/L CIP with a reactor

volume of 500 mL [49]. In this study, the electrical energy per order of CIP degraded

by SnO2-Sb/Ti electrode was 11.81 Wh/L for an initial concentration of 50 mg/L CIP

(150 M) with a volume of 250 mL. The results indicated that the degradation of the

CIP by electrochemical oxidation had a lower energy cost.

Fig. 6(a) shows the relationship between the reaction rate constant (k) and the

EE/O at different applied current densities. When the applied current density

increased, the EE/O increased. However, the evaluation of the system depends not

only on the energy cost but also on the space efficiency, which is related to the

reaction rate constant (k). A good system requires not only a low energy demand but

also high space efficiency. In Fig. 6(a), the dashed line represents the fact that one unit

of k value increases when one unit of EE/O was consumed. When the applied current

density increased from 10 to 30 mA/cm2, the increase in the k value was clearly

higher than that of EE/O. That is, the EE/O at 30 mA/cm2 was 1.9 times as high as

that at 10 mA/cm2, while the k value at 30 mA/cm2 was 2.1 times as high as that at 10

mA/cm2. However, when the applied current density increased from 30 to 40 mA/cm2,

22
the k value decreased while the EE/O increased quickly. Therefore, the applied

current density of 30 mA/cm2 was the most suitable.

As seen from Fig. 6(b), when the initial concentration increased from 10 to 100

mg/L, the EE/O increased gradually from 5.98 to 31.61 Wh/L. In addition to EE/O,

one of the evaluation in a practical application is the energy efficiency. However, very

little research has been conducted on EE/O combined with energy efficiency. To

comprehensively evaluate the electrochemical degradation process, it is of great

significance to not only consider the EE/O but also the energy efficiency. The plot of

energy efficiency versus initial concentration showed that the highest energy

efficiency was observed at the initial concentration of 50 mg/L. When the initial

concentration was further increased, the energy efficiency decreased instead. When

the initial concentration of CIP was lower than 50 mg/L, the chance of contact

between the pollutants and the hydroxyl radicals increased with a rise in initial

concentration. Thus the energy efficiency increases. However, when the initial

concentration was 100 mg/L, many intermediates were accumulated (Fig. 3(c)). These

intermediates could compete with the degradation of CIP and thus decrease the energy

efficiency.

The change in the EE/O at different pH values is shown in Fig 6(c). Generally,

there is no significant difference for EE/O at different pH values. This result revealed

that pH had little effect on EE/O for CIP degradation using a SnO2-Sb/Ti electrode.

This result was in accordance with that obtained in section 3.2.4.

4. Conclusions

23
This study focused on the effects of current density, the initial concentration of

CIP and the initial pH on the electrochemical oxidation of CIP using a SnO2-Sb/Ti

electrode. The results demonstrated that CIP can be effectively degraded by

electrochemical oxidation. After 120 min of electrolysis, the removal of CIP (50

mg/L), COD and TOC at a current density of 30 mA/cm2 was almost 99.5%, 86.0%

and 70.0%, respectively. With an increase in the applied current density (1040

mA/cm2), the CIP and COD degradation efficiency first increased and then decreased;

the ACE decreased at a higher current density. The reaction rate constant (k) first

increased with the rise in EE/O and then decreased at an applied current density of 40

mA/cm2, even though the EE/O was further enhanced. Considering the energy

efficiency and space efficiency, the applied current density of 30 mA/cm 2 was the

most suitable. Higher initial CIP concentrations inhibited the removal efficiency of

CIP degradation, but the absolute amount of CIP removal increased. The inhibition of

CIP removal efficiency may be because more intermediates were produced at the

higher CIP concentration and these intermediates competed with CIP for the active

sites. An increase of the absolute amount of CIP removal was attributed to a higher

initial CIP concentration that enhanced its concentration gradient and mass transfer

across the diffusion layer. Higher initial CIP concentrations resulted in greater ACE

be a e he a al l ea ap ly w h gan ma e bef e xygen wa

evolved. The highest energy efficiency was observed at an initial concentration of 50

mg/L. The initial pH had no significant effect. Under different conditions the CIP

degradation was in good agreement with an apparent first-order kinetic model. The

24
electrochemical degradation pathways of CIP in aqueous solution were proposed as

the oxidative degradation of the piperazinyl moiety, hydroxylation, and defluorination.

Inorganic N compounds were and . F was reduced to F-.

Acknowledgments

This work was supported by the National Natural Science Foundation of China

(No. 51578070) and Special Project of National International Cooperation in Science

and Technology of china (No.2013DFR90290).

25
References
[1] F. Yuan, C. Hu, X. X. Hu, D. B. Wei, Y. Chen, J. H. Qu, Photodegradation and toxicity
changes of antibiotics in UV and UV/H2O2 process, J. Hazard. Mater. 185 (2011) 1256-1263.
[2] C. D. Qi, X. T. Liu, C. Y. Lin, X. H. Zhang, J. Ma, H. B. Tan, W. Ye, Degradation of
sulfamethoxazole by microwave-activated persulfate: Kinetics, mechanism and acute toxicity,
Chem. Eng. J. 249 (2014) 6-14.
[3] X. V. Doorslaer, K. Demeestere, P. M. Heynderickx, H. V. Langenhove, J. Dewulf, UV-A
and UV-C induced photolytic and photocatalytic degradation of aqueous ciprofloxacin and
moxifloxacin: reaction kinetics and role of adsorption, Appl. Catal. B-environ. 101 (2011)
540-547.
[4] A. Jia, Y. Wan, Y. Xiao, J. Y. Hu, Occurrence and fate of quinolone and fluoroquinolone
antibiotics in a municipal sewage treatment plant, Water. Res. 46 (2012) 387-394.
[5] H. Yao, P. Z. Sun, D. Minakata, J. C. Crittenden, C. H. Huang, Kinetics and Modeling of
Degradation of Ionophore Antibiotics by UV and UV/H2O2, Environ. Sci. Technol. 47 (2013)
4581-4589.
[6] I. Sirs, E. Brillas, M. A. Oturan, M. A. Rodrigo, M. Panizza, Electrochemical advanced
oxidation processes: today and tomorrow. A review, Environ. Sci. Pollut. Res. 21 (2014)
8836-8367.
[7] I. Sirs, E. Brillas, Remediation of water pollution caused by pharmaceutical residues based
on electrochemical separation and degradation technologies: A review, Environ. Int. 40 (2012)
212-229.
[8] M. H. Zhou, Q. Z. Dai, L. C. Lei, C. A. Ma, D. H. Wang, Long life modified lead dioxide
anode for organic wastewater treatment: electrochemical characteristics and degradation
mechanism, Environ. Sci. Technol. 39 (2005) 363-370.
[9] F. Sopaj, M. A. Rodrigo, N. Oturan, F. I. Podvorica, J. Pinson, M. A. Oturan, Influence of
the anode materials on the electrochemical oxidation efficiency. Application to oxidative
degradation of the pharmaceutical amoxicillin, Chem. Eng. J. 262 (2015) 286-294.
[10] A. Fab a ka, A. B. B el ka, P. Stepnowski, S. Stolte, E. M. Siedlecka, Electrochemical
degradation of sulfonamides at BDD electrode: Kinetics, reaction pathway and eco-toxicity
evaluation; J. Hazard. Mater. 280 (2014) 579-587.
[11] J. Wu, H. Zhang, N. Oturan, Y. Wang,L. Chen, M. A. Oturan, Application of response
surface methodology to the removal of the antibiotic tetracycline by electrochemical process
using carbon-felt cathode and DSA (Ti/RuO2IrO2) anode, Chemosphere. 87 (2012) 614-620.
[12] X. Xiao, X. Zeng, A. T. Lemley, Species-dependent degradation of ciprofloxacin in a
membrane anodic Fenton system. J. Agr. Food. Chem. 58 (2010) 10169-10175.
[13] T. S. Chen, Y. M. Kuo, J. L. Chen, K. L. Huang, Anodic Degradation of Ofloxacin on a
Boron-Doped Diamond Electrode, Int. J. Electrochem. Sci. 8 (2013) 7625-7633.
[14] E. Guinea, J. A. Garrido, R. M. Rodrguez, P. L. Cabot, C. Arias, F. Centellas, E. Brillas,
Degradation of the fluoroquinoloneenrofloxacin by electrochemical advanced oxidation
processes based on hydrogen peroxide electrogeneration, Electrochim. Acta. 55 (2010)
2101-2115.
[15] E. Guinea, E. Brillas, F. Centellas, P. Caizares, M. A. Rodrigo, C. Sez, Oxidation of
enrofloxacin with conductive-diamond electrochemical oxidation, ozonation and Fenton

26
oxidation. A comparison, Water. Res. 43 (2009) 2131-2138.
[16] V. S. Antonin, M. C. Santos, S. Garcia-Segura, E. Brillas, Electrochemical incineration of
the antibiotic ciprofloxacin in sulfate medium and synthetic urine matrix, Water. Res. 83 (2015)
31-41.
[17] H. Lin, J. F. Niu, J. L Xu, Y. Li, Y. H. Pan, Electrochemical mineralization of
sulfamethoxazole by Ti/SnO2-Sb/Ce-PbO2 anode: Kinetics, reaction pathways, and energy cost
evolution, Electrochim. Acta. 97 (2013) 167-174.
[18] A. El-Ghenymy, J. A. Garrido, F. Centellas, C. Arias, P. L. Cabot, R. M. Rodrguez, E.
Brillas, Electro-Fenton and photoelectro-Fenton degradation of sulfanilicacid using a
boron-doped diamond anode and an air diffusion cathode, J. Phys. Chem. A. 116 (2012)
3404-3412.
[19] J. F. Niu , Y. P. Bao, Y. Li , Z. Chai, Electrochemical mineralization of pentachlorophenol
(PCP) by Ti/SnO2-Sb electrodes, Chemosphere. 92 (2013) 1571-1577.
[20] J. Q. Fan, G. H. Zhao, H. Y. Zhao, S. N. Chai, T. C. Cao, Fabrication and application of
mesoporous Sb-doped SnO2 electrode with high specific surface in electrochemical degradation
of ketoprofen, Electrochim. Acta. 94 (2013) 21-29.
[21] J. Radjenovic, B. I. Escher, K. Rabaey, Ele hem al eg a a n f he -blocker
metoprolol by Ti/Ru0.7Ir0.3O2 and Ti/SnO2-Sb electrodes, Water. Res. 45 (2011) 3025-3214.
[22] A. Y. Bagastyo, D. J. Batstone, K. Rabaey, J. Radjenovic, Electrochemical oxidation of
electrodialysed reverse osmosis concentrate on Ti/Pt-IrO2, Ti/SnO2-Sb and boron-doped
diamond electrodes, Water. Res. 47 (2013) 242-250.
[23] T. G. Duan, Q. Wen, Y. Chen, Y. D. Zhou, Y. Duan, Enhancing electrocatalytic
performance of Sb-doped SnO2 electrode by compositing nitrogen-doped graphene nanosheets,
J. Hazard. Mater. 280 (2014) 304-314.
[24] D. G. J. Larsson, C. D. Pedro, N. Paxeus, Effluent from drug manufactures contains
extremely high levels of pharmaceuticals, J. Hazard. Mater. 148 (2007) 751-755.
[25] H. Lin, J. F. Niu, S. Y. Ding, L. L. Zhang, Electrochemical degradation of
perfluorooctanoic acid (PFOA) by Ti/SnO2-Sb, Ti/SnO2-Sb/PbO2 and Ti/SnO2-Sb/MnO2
anodes, Water. Res. 46 (2012) 2281-2289.
[26] B. C. Lozano, C. Comninellis, A. D. Battisti, Service life of Ti/SnO2-Sb2O5 anodes, J. Appl.
Electrochem. 27 (1997) 970-974.
[27] Y. Wang, Z. Y. Shen, X. C. Chen, Effects of experimental parameters on 2,4-dichlorphenol
degradation over Er-chitosan-PbO2 electrode, J. Hazard. Mater. 178 (2010) 867874.
[28] Y. Bessekhouad, R. Brahimia, F. Hamdini, M. Trari, Cu2S/TiO2 heterojunction applied to
visible light Orange II degradation, J. Photoch. Photobio. A. 48 (2012) 15-23.
[29] W. Qiu, Y. Zheng, K. A. Haralampides, Study on a novel POM-based magnetic
photocatalyst: Photocatalytic degradation and magnetic separation, Chem. Eng. J. 125, (2007)
165-176.
[30] L. C. Zhang, L. Xu, J. He, J. J. Zhang, Preparation of Ti/SnO2-Sb electrodes modified by
carbon nanotube for anodic oxidation of dye wastewater and combination with nanofiltration,
Electrochim. Acta. 117 (2014) 192-201.
[31] China EPA, Monitoring and analysis method for water and waste water, fourth ed., China
Environmental Science Press, Beijing, 2002, pp 232-235.
[32] H. S. Lee, H. J. Lee, J. Jeong, J. Lee, N. B. Park, C. Lee, Activation of persulfates by

27
carbon nanotubes: Oxidation of organic compounds by nonradical mechanism, Chem. Eng. J.
266 (2015) 28-33.
[33] M. Q. He, W. B. Li, J. X. Xia, L. Xu, J. Di, H. Xu, S. Yin, H. M. Li , M. N. Li, The
enhanced visible light photocatalytic activity of yttrium-doped BiOBr synthesized via a
reactable ionic liquid, Appl. Surf. Sci. 331 (2015) 170-178.
[34] D. Wu, P. W. Huo, Z. Y. Lua, X. Gao, X. L. Liu, W. D. Shi, Y. S. Yan, Preparation of
heteropolyacid/TiO2/fly-ash-cenosphere photocatalyst for the degradation of ciprofloxacin from
aqueous solutions, Appl. Surf. Sci. 258 (2012) 7008-7015.
[35] M. S. Yahya, N. Oturan, K. E. Kacemi, M. E. Karbane, C. T. Aravindakumar, M. A. Oturan,
Oxidative degradation study on antimicrobial agent ciprofloxacin by electro-fenton process:
Kinetics and oxidation products, Chemosphere, 117 (2014) 447-454.
[36] Y. Chen, L. Hong, W. Q. Han, L. J. Wang, X. Y. Sun, J. S. Li, Treatment of high explosive
production wastewater containing RDX by combined electrocatalytic reaction and anoxic-oxic
biodegradation, Chem. Eng. J. 168 (2011) 1256-1262.
[37] S. Song, J. Q. Fan, Z. Q. He, L. Y. Zhan, Z. W. Liu, J. M. Chen, X. H. Xu, Electrochemical
degradation of azo dye C.I. Reactive Red 195 by anodic oxidation on Ti/SnO2-Sb/PbO2
electrodes, Electrochim. Acta. 55 (2010) 3606-3613.
[38] H. Li, Q. N. Yu, B. Yang, Z. J. Li, L.C. Lei, Electrochemical treatment of artificial
humidity condensate by large-scale boron doped diamond electrode, Sep. Purif. Technol. 138
(2014) 13-20.
[39] S. Bab , M. e a, . k , h ly eg a a n f n fl xa n, en fl xa n an
ciprofloxacin in various aqueous media, Chemosphere. 91 (2013) 1635-1642.
[40] X. L. Liu, P. Lv, G. X. Yao, C. C. Ma, Y. F. Tang, Y. T. Wu, P. W. Huo, J. M. Pan, W. D. Shi,
Y. S. Yan, Selective degradation of ciprofloxacin with modified NaCl/TiO2 photocatalyst by
surface molecular imprinted technology, Colloids and Surfaces A: Physicochem. Eng. Aspects
441 (2014) 420-426.
[41] M. E. Kemary, H. E. Shamy, I. E. Mehasseb, Photocatalytic degradation of ciprofloxacin
drug in water using ZnO nanoparticles, J. Lumin. 130 (2010) 2327-2331.
[42] L. Xu, Z. Guo, L. S. Du, J. He, Decolourization and degradation of C.I. Acid Red 73 by
anodic oxidation and the synergy technology of anodic oxidation coupling nanofiltration,
Electrochim. Acta. 97 (2013) 150-159.
[43] J. F. Niu, D. Maharana, J. L. Xu, Z. Chai, Y. P. Bao, A high activity of Ti/SnO2-Sb
electrode in the electrochemical degradation of 2, 4-dichlorophenol in aqueous solution, J.
Environ. Sci- China. 25 (2013) 1424-1430.
[44] Y. Ji, C. Ferronato, A. Salvador, X. Yang, J. M. Chovelon, Degradation of ciprofloxacin
and sulfamethoxazole by ferrous-activated persulfate: implications for remediation of
groundwater contaminated by antibiotics, Sci. Total Environ 472 (2014) 800-808.
[45] X. X. Zhang, R. P. Li, M. K. Jia, S. L. Wang, Y. P. Huang, Degradation of ciprofloxacin in
aqueous bismuth oxybromide (BiOBr) suspensions under visible light irradiation: A direct hole
oxidation pathway, Chem. Eng. J. 274 (2015) 290-297.
[46] H. G. Guo, N. Y. Gao, W. H. Chu, L. Li, Y. J. Zhang, J. S. Gu, Y. L. Gu, Photochemical
degradation of ciprofloxacin in UV and UV/H2O2 process: kinetics, parameters, and products,
Environ. Sci. Pollut. R. 20 (2013) 3202-3213.
[47] T. C. An, H. Yang, G. Y. Li, W. H. Song, W. J. Cooper, X. P. Nie, Kinetics and mechanism

28
of advanced oxidation processes (AOPs) in degradation of ciprofloxacin in water, APPL. Catal.
B-Environ. 94 (2010) 288-294.
[48] E. Kugelmann, C. R. Albert, G. Bringmann, U. lzg abe, Fen n x a n: A lf
the investigation of potential drug metabolites, J. Pharmaceut. Biomed. 54 (2011) 1047-1058.
[49] M. M. Ahmed, S. Chiron, Ciprofloxacin oxidation by UV-C activated peroxymonosulfate
in wastewater, J. Hazard. Mater. 265 (2014) 41-46.

29
Legends of Figures and Tables

Fig. 1 (a), (b) SEM images and (c) XRD patterns of the SnO2-Sb/Ti electrode.

Fig. 2 (a) CV curves of SnO2-Sb/Ti electrode in 0.5 mol/L Na2SO4 in the absence of

CIP and in the presence of 50mg/L CIP at the scan rate of 10 mV/s. (b) CIP, COD and

TOC removal efficiency of the system. (c) Accelerated service life test on SnO2-Sb/Ti

electrode in 0.5 mol/L H2SO4 with a current density of 100 mA/cm2.

Fig. 3 (a) Effect of the applied current density on CIP and COD removal efficiency;

Inset: the correlation between Ln (CIP0/CIPt) and Ln (COD0/CODt) versus

electrolysis time (initial CIP concentration, 50 mg/L; initial pH, 5.4; T, 30 C; 0.05

mol/L Na2SO4; plate distance, 20 mm); (b) Effect of the initial CIP concentration on

CIP and COD removal efficiency (applied current density, 30 mA/cm2; initial pH, 5.4;

T, 30 C; 0.05 mol/L Na2SO4; plate distance, 20 mm); (c) Change of UV absorption

spectra of 30 and 100 mg/L CIP solution (applied current density, 30 mA/cm2; initial

pH, 5.4; T, 30 C; 0.05 mol/L Na2SO4; plate distance, 20 mm); (d) Effect of the initial

pH concentration on CIP and COD removal efficiency (applied current density, 30

mA/cm2; initial CIP concentration, 50 mg/L; T, 30 C; 0.05 mol/L Na2SO4; plate

distance, 20 mm).

Fig. 4 Evolution of the inorganic intermediates during electrolysis (initial CIP

concentration, 50 mg/L; initial pH, 5.4; T, 30 C; 0.05 mol/L Na2SO4; plate distance,

20 mm).
30
Fig. 5 The proposed degradation pathway for CIP in an anodic oxidation system

(initial CIP concentration, 50 mg/L; initial pH, 5.4; T, 30 C; 0.05 mol/L Na2SO4;

plate distance, 20 mm).

Fig. 6 (a) The trend between the apparent rate constant and the electrical energy

consumption per order of CIP destruction; Inset: The EE/O at different current density

(initial CIP concentration, 50mg/L; initial pH, 5.4; T, 30 C; 0.05 mol/L Na2SO4; plate

distance, 20 mm); (b) the trend between the energy efficiency and the initial CIP

concentration; Inset: The EE/O at different CIP concentration (applied current density,

30 mA/cm2; initial pH, 5.4; T, 30 C; 0.05 mol/L Na2SO4; plate distance, 20 mm); (c)

the electrical energy consumption per order of CIP destruction at different pH values

(applied current density, 30 mA/cm2; initial CIP concentration, 50 mg/L; T, 30 C;

0.05 mol/L Na2SO4; plate distance, 20 mm).

Tab. 1 The kinetics for CIP and COD removal by SnO2-Sb/Ti anode.

Tab. 2 Identification of the CIP degradation products by HPLCMS.

31
(a) (b)

(c)

Fig. 1 Ying Wang et al.

32
Fig. 2 Ying Wang et al.

33
34
Fig. 3 Ying Wang et al.

35
Fig. 4 Ying Wang et al.

36
Fig. 5 Ying Wang et al.

37
Fig. 6 Ying Wang et al.

38
Table. 1 The kinetics for CIP and COD removal by SnO2-Sb/Ti anode

Initial-final Average
Parameters pH during voltage kCIP R2 kCOD R2 ACE
reaction V (min-1) (min-1)
(a) Current density (mA/cm2)
10 5.4-4.24 6.1 0.022 0.976 0.008 0.991 0.130
20 5.4-4.21 7.2 0.033 0.979 0.009 0.982 0.073
30 5.4-4.74 7.8 0.046 0.991 0.017 0.999 0.063
40 5.4-4.64 8.5 0.039 0.992 0.010 0.976 0.038
(b) Initial concentration (C0, mg/L)
10 5.4-4.86 7.8 0.090 0.995 0.031 0.970 0.015
30 5.4-4.47 7.8 0.056 0.989 0.024 0.991 0.042
50 5.4-4.74 7.9 0.046 0.991 0.017 0.999 0.063
100 5.4-4.58 8.1 0.018 0.991 0.007 0.978 0.082
(c) Initial pH
3 3-2.80 7.7 0.051 0.978 0.018 0.979 0.063
5 5-4.47 7.7 0.045 0.996 0.016 0.989 0.061
7 7-5.69 7.7 0.045 0.995 0.014 0.994 0.058
9 9-6.14 7.7 0.043 0.989 0.011 0.998 0.053
11 11-7.80 7.7 0.048 0.992 0.013 0.961 0.057

Tab. 1 Ying Wang et al.

39
Table. 2 Identification of the CIP degradation products by HPLCMS.

[M+H]+ tr(min) Identification Chemical structure

332 2.30 C17H18FN3O3

330 1.28 C17H19N3O4

344 1.59 C17H17N3O5

291 3.15 C14H12FN2O4

263 3.25 C13H11FN2O3

348 2.48 C17H18FN3O4

366 2.01 C17H20FN3O5

Tab. 2 Ying Wang et al.

40
SnO2-Sb/Ti electrode efficiently degraded CIP with low energy in a wide pH

range.

Current density and initial CIP concentration greatly affected the process.

Oxidation routes were piperazine ring cleavage, hydroxylation and

defluorination.

The process was evaluated by energy demand combined with energy

efficiency.

Das könnte Ihnen auch gefallen