7
Robust TiO 2 /BDD heterojunction photoanodes for determination of chemical oxygen demand in wastewatersYanhe Han, a Jingxia Qiu, ab Yuqing Miao, b Jisheng Han, c Shanqing Zhang, * a Haimin Zhang a and Huijun Zhao a Received 31st March 2011, Accepted 26th June 2011 DOI: 10.1039/c1ay05193h A TiO 2 /BDD heterojunction photoanode, utilizing the inherent properties of nanostructured titanium dioxide (TiO 2 ) and boron-doped diamond (BDD), was prepared and used to determine chemical oxygen demand (COD) in wastewaters. The TiO 2 nanoparticles were dip-coated on a BDD substrate and subject to calcination processes. A uniform, continuous and robust mixed-phase (anatase and rutile) TiO 2 /BDD heterojunction electrode was obtained. The TiO 2 /BDD heterojunction electrode was evaluated using a series of materials characterisation, electrical and electrochemical techniques. The preliminary results suggest the elevated photoelectrocatalytic activity over the oxidation of organic compounds stemmed from the formation of the p–n junction of the TiO 2 /BDD electrode. The TiO 2 / BDD electrode has an excellent resistance towards strong acid due to the use of BDD substrate, which is an added advantage for practical application. Under the optimized experimental conditions, the TiO 2 /BDD electrode is capable of indiscriminately oxidizing a wide spectrum of organic compounds in a photoelectrochemical thin-layer cell. This bestows the photoelectrochemical system with the ability to measure the COD of synthetic and real samples in a fast, sensitive, reproducible and accurate fashion. In particular, a typical analysis time of 5 minutes, a practical detection limit of 0.12 mg L 1 COD, a RSD% value of 1.5% and a linear range of 0–300 mg L 1 were achieved. The TiO 2 /BDD electrode can be an ideal sensor for online and in situ monitoring of organic pollutants in wastewaters. 1. Introduction TiO 2 has been widely used in the fields of water purification/ sterilisation, 1 wastewater treatment 2 and water quality moni- toring, 3 owing to its outstanding photocatalytic activity, excel- lent chemical and photochemical stability, superior oxidation ability, low cost, and environmentally friendly performance. However, the quantum efficiency was still very low during the photocatalytic processes mainly due to the fast recombination of photoholes and photoelectrons generated under the illumination of UV light. In order to enhance the oxidation efficiency of photocatalysis at the TiO 2 surface, TiO 2 was immobilised or grown on various conducting substrates, such as ITO 4 and tita- nium metals, 5 to allow the application of potential, which was an effective way to significantly diminish the recombination rates. It is well recognised that boron-doped diamond (BDD) is one of the most promising electrode materials for this century. 6–9 Due to the characteristics of small background current, low noise signals, wide potential window, high mechanical strength and resilient corrosion resistance (even being anodic polarised in acidic solutions), 10–13 and a durable resistance to surface fouling by impurities and long term response stability, BDD is widely used as an electrode or electrode substrate combining with other functional materials. Moreover, because BDD exhibits p-type semiconductor characteristics, 14 numerous attempts to prepare TiO 2 /BDD heterojunction photoanode systems with n-type TiO 2 have been successful. 15–17 It was demonstrated that these heter- ojunction electrodes possessed improved oxidation power to organic compounds in wastewater. Chemical oxygen demand (COD) is one of the most important parameters and has been widely employed for water quality assessment. The standard method (potassium dichromate method) for COD determination involves a 2 hour hydrothermal digestion process under high pressure and temperature. It also requires the use of expensive (Ag 2 SO 4 ), highly corrosive (H 2 SO 4 ) and toxic (Cr 2 O 7 and HgSO 4 ) reagents. 18 Consequently, the secondary pollution is unavoidable when the standard method is employed. Numerous electrochemical sensors such as copper electrode, 19 Cu/CuO electrode, 20 nano-PbO 2 modified electrode, 21 F-doped a Environmental Futures Centre and Griffith School of Environment, Gold Coast Campus, Griffith University, QLD, 4222, Australia. E-mail: s.zhang@griffith.edu.au; Fax: +61-7-5552 8067; Tel: +61-7-5552 8155 b Institute of Physical Chemistry, Zhejiang Normal University, Jinhua, Zhejiang Province, 321004, P.R. China c QLD Micro- and Nanotechnology Centre, Nathan Campus, Griffith University, QLD, 4111, Australia † Electronic supplementary information (ESI) available: supporting images. See DOI: 10.1039/c1ay05193h This journal is ª The Royal Society of Chemistry 2011 Anal. Methods, 2011, 3, 2003–2009 | 2003 Dynamic Article Links C < Analytical Methods Cite this: Anal. Methods, 2011, 3, 2003 www.rsc.org/methods PAPER Published on 01 August 2011. Downloaded by Imperial College London Library on 21/05/2013 04:54:31. View Article Online / Journal Homepage / Table of Contents for this issue

Robust TiO2/BDD heterojunction photoanodes for determination of chemical oxygen demand in wastewaters

  • Upload
    huijun

  • View
    214

  • Download
    0

Embed Size (px)

Citation preview

Dynamic Article LinksC<AnalyticalMethods

Cite this: Anal. Methods, 2011, 3, 2003

www.rsc.org/methods PAPER

Publ

ishe

d on

01

Aug

ust 2

011.

Dow

nloa

ded

by I

mpe

rial

Col

lege

Lon

don

Lib

rary

on

21/0

5/20

13 0

4:54

:31.

View Article Online / Journal Homepage / Table of Contents for this issue

Robust TiO2/BDD heterojunction photoanodes for determination of chemicaloxygen demand in wastewaters†

Yanhe Han,a Jingxia Qiu,ab YuqingMiao,b Jisheng Han,c Shanqing Zhang,*a Haimin Zhanga and Huijun Zhaoa

Received 31st March 2011, Accepted 26th June 2011

DOI: 10.1039/c1ay05193h

A TiO2/BDD heterojunction photoanode, utilizing the inherent properties of nanostructured titanium

dioxide (TiO2) and boron-doped diamond (BDD), was prepared and used to determine chemical

oxygen demand (COD) in wastewaters. The TiO2 nanoparticles were dip-coated on a BDD substrate

and subject to calcination processes. A uniform, continuous and robust mixed-phase (anatase and

rutile) TiO2/BDD heterojunction electrode was obtained. The TiO2/BDD heterojunction electrode was

evaluated using a series of materials characterisation, electrical and electrochemical techniques. The

preliminary results suggest the elevated photoelectrocatalytic activity over the oxidation of organic

compounds stemmed from the formation of the p–n junction of the TiO2/BDD electrode. The TiO2/

BDD electrode has an excellent resistance towards strong acid due to the use of BDD substrate, which

is an added advantage for practical application. Under the optimized experimental conditions, the

TiO2/BDD electrode is capable of indiscriminately oxidizing a wide spectrum of organic compounds in

a photoelectrochemical thin-layer cell. This bestows the photoelectrochemical system with the ability to

measure the COD of synthetic and real samples in a fast, sensitive, reproducible and accurate fashion.

In particular, a typical analysis time of 5 minutes, a practical detection limit of 0.12 mg L�1 COD,

a RSD% value of 1.5% and a linear range of 0–300 mg L�1 were achieved. The TiO2/BDD electrode can

be an ideal sensor for online and in situ monitoring of organic pollutants in wastewaters.

1. Introduction

TiO2 has been widely used in the fields of water purification/

sterilisation,1 wastewater treatment2 and water quality moni-

toring,3 owing to its outstanding photocatalytic activity, excel-

lent chemical and photochemical stability, superior oxidation

ability, low cost, and environmentally friendly performance.

However, the quantum efficiency was still very low during the

photocatalytic processes mainly due to the fast recombination of

photoholes and photoelectrons generated under the illumination

of UV light. In order to enhance the oxidation efficiency of

photocatalysis at the TiO2 surface, TiO2 was immobilised or

grown on various conducting substrates, such as ITO4 and tita-

nium metals,5 to allow the application of potential, which was an

effective way to significantly diminish the recombination rates.

aEnvironmental Futures Centre and Griffith School of Environment, GoldCoast Campus, Griffith University, QLD, 4222, Australia. E-mail:[email protected]; Fax: +61-7-5552 8067; Tel: +61-7-5552 8155bInstitute of Physical Chemistry, Zhejiang Normal University, Jinhua,Zhejiang Province, 321004, P.R. ChinacQLD Micro- and Nanotechnology Centre, Nathan Campus, GriffithUniversity, QLD, 4111, Australia

† Electronic supplementary information (ESI) available: supportingimages. See DOI: 10.1039/c1ay05193h

This journal is ª The Royal Society of Chemistry 2011

It is well recognised that boron-doped diamond (BDD) is one

of the most promising electrode materials for this century.6–9 Due

to the characteristics of small background current, low noise

signals, wide potential window, high mechanical strength and

resilient corrosion resistance (even being anodic polarised in

acidic solutions),10–13 and a durable resistance to surface fouling

by impurities and long term response stability, BDD is widely

used as an electrode or electrode substrate combining with other

functional materials. Moreover, because BDD exhibits p-type

semiconductor characteristics,14 numerous attempts to prepare

TiO2/BDD heterojunction photoanode systems with n-type TiO2

have been successful.15–17 It was demonstrated that these heter-

ojunction electrodes possessed improved oxidation power to

organic compounds in wastewater.

Chemical oxygen demand (COD) is one of the most important

parameters and has been widely employed for water quality

assessment. The standard method (potassium dichromate

method) for COD determination involves a 2 hour hydrothermal

digestion process under high pressure and temperature. It also

requires the use of expensive (Ag2SO4), highly corrosive (H2SO4)

and toxic (Cr2O7� and HgSO4) reagents.18 Consequently, the

secondary pollution is unavoidable when the standard method is

employed.

Numerous electrochemical sensors such as copper electrode,19

Cu/CuO electrode,20 nano-PbO2 modified electrode,21 F-doped

Anal. Methods, 2011, 3, 2003–2009 | 2003

Publ

ishe

d on

01

Aug

ust 2

011.

Dow

nloa

ded

by I

mpe

rial

Col

lege

Lon

don

Lib

rary

on

21/0

5/20

13 0

4:54

:31.

View Article Online

PbO2 modified electrode,22 RhO3/Ti electrode,23 boron-doped

diamond electrode,24 and rotating Pt ring-Pt/PbO2 disc elec-

trode25 have been applied for the determination of COD. The

main advantages of electrochemical methods for COD analysis

were simplicity, short response time, wide linear range, low cost,

and the ease to be automated. However, insufficient oxidation

capability, instable background current of the electrodes and the

existence of persistent organic pollutants in wastewater lead to

problems of poor accuracy and reproducibility.

Recently, a very simple COD analytical method using TiO2

electrodes for the photoelectrocatalytic process, namely PeCOD,

was developed by our group.4,26,27 The PeCOD method can

effectively tackle the above problems of the standard method,

and in fact, it was also achieved by using other TiO2 electrodes,

such as TiO2/Ti electrodes.28 However, these electrodes still need

to be improved in terms of photocatalytic activity and stabilities

in acidic environments.16,29 The TiO2/BDD heterojunction pho-

toanode is a promising candidate to be used for this PeCOD

technology for online and in situ applications, where the

robustness of the sensor is an essential criterion.

In this work, TiO2/BDD electrodes were fabricated by dip-

coating a TiO2 layer onto BDD electrode surfaces and subsequent

calcination processes. The as prepared TiO2/BDD electrodes were

characterized using a series of materials characterisation, elec-

trical and electrochemical techniques. The photoelectrocatalytic

activity over the oxidation of various organic compounds was

measured and optimized in a photoelectrochemical thin-layer cell.

Under the optimized experimental conditions, the TiO2/BDD

electrodes were used to detect the COD concentration of

numerous synthetic and real samples to validate the suitability in

PeCOD application.

2. Experimental

2.1. Materials and chemicals

BDD electrodes with a resistivity of 0.1 U cm were purchased

from CSEM (Switzerland). Indium tin oxide (ITO) conducting

glass slides (8 U per square) were supplied by Delta Technologies

Limited. Titanium butoxide (97%, Aldrich), potassium hydrogen

phthalate (AR, Aldrich), D-glucose (99%, BDH), and sodium

nitrate (AR, Aldrich) were used as received. All other chemicals

used in this work were of analytical reagent and used as received

without further purification. All solutions were prepared using

high purity deionized water (>18 MU cm; Millipore Corp.).

2.2. Preparation of the TiO2/BDD electrodes

Aqueous TiO2 colloid was prepared by hydrolysis of titanium

butoxide according to the method described in our previous

work.30 Briefly, a mixture of 25 mL titanium butoxide and 8 mL

propan-2-ol was added, dropwise and at room temperature, to

300 mL of a 0.1 M nitric acid solution under vigorous stirring,

resulting in a slurry. The slurry was heated to 80 �C and stirred

vigorously for 10 h to achieve peptization. The resulting colloid

was then hydrothermally treated in an autoclave at 200 �C for 12

h. The colloidal suspension was then introduced into a rotary

evaporator and evaporated to a final solid concentration of ca.

6% w/v with particle sizes ranging from 8 to 10 nm. Carbowax

(30 wt% with respect to the mass of TiO2 solids) was added to

2004 | Anal. Methods, 2011, 3, 2003–2009

increase the porosity and reinforce the mechanical strength of the

TiO2 thin-film. The BDD electrode was pre-treated with aqua

regia and H2O2 solution at 70 �C, each for 30 min, and then

rinsed with distilled water to remove the surface oxides formed

during the acid pretreatment.31 After the pretreatment, the BDD

slides were dip-coated in the TiO2 colloidal solution with

a withdrawing speed of 2 mm s�1. The coated electrodes were

calcined at 450 �C in air for 0.5 h and 700 �C in Ar protective

atmosphere. For comparison, TiO2/ITO electrodes were fabri-

cated as a control using the same procedures except different

substrates.

2.3. Apparatus and measurements

2.3.1. Materials characterisation. The surface morphology of

the TiO2 samples was observed using SEM on the JEOL JSM-

6300 field emission scanning electron microscopy (FESEM,

Tokyo, Japan). X-Ray diffraction (XRD) spectra were performed

with a Philips PW1050 diffractometer using CuKa radiation.

2.3.2. Direct electrical measurement. The current density (J)

–voltage relationships of the TiO2/BDD electrodes were obtained

by using a Keithley 4200 Semiconductor Characterization System

and amicromanipulator station in air at room temperature (23 �C).

2.3.3. Photoelectrochemical characterization. Electro-

chemical experiments, i.e., linear scanning voltammetry (LSV),

cyclic voltammetry (CV) and chronoamperometry at

a controlled potential (+0.4 V vs. Ag/AgCl electrode), were

performed under UV illumination (6.6 mW cm�2 at 365 nm) at

the room temperature in a three-electrode photoelectrochemical

bulk cell with a quartz window for UV illumination (see

Fig. S1†).32 The working electrode surface area exposed to

solution was a circle with a diameter of 6 mm, which was the

photoelectrochemical reaction area. A Ag/AgCl reference elec-

trode and platinummesh were used as the reference and auxiliary

electrodes, respectively. Illumination was achieved with a 150 W

xenon arc lamp light source with quartz focusing lenses (HF-200

w-95). To minimize sample heating from the infrared fraction of

the light source, the beam was passed through an UV-band pass

filter (UG-5, Schott) prior to illuminating the electrode surface.

The light intensity was measured with a UV-irradiance meter

(UVA, Instruments of Beijing Normal University). A voltam-

mograph (CV-27, BAS) was used as a potentiostat for the elec-

trochemical measurements.

2.3.4. PeCOD measurement. The PeCOD measurements

were carried out in a photoelectrochemical thin-layer cell reactor

using a stopped-flow technique (see Fig. S2†).4 The volume of the

thin-layer reactor in this work was 2.4 mL, the applied potential

was +0.40 V and the intensity of the illumination was 6.6 mW

cm�2. In order to ensure sufficient conductivity in the thin-layer

electrochemical system, a certain amount of solid NaNO3

equivalent to 2 M was added to the tested samples.

2.4. Calculation of PeCOD values

The general equation for mineralization on a TiO2 electrode in

the thin-layer cell can be summarised as follows:4

This journal is ª The Royal Society of Chemistry 2011

Publ

ishe

d on

01

Aug

ust 2

011.

Dow

nloa

ded

by I

mpe

rial

Col

lege

Lon

don

Lib

rary

on

21/0

5/20

13 0

4:54

:31.

View Article Online

CyHmOjNkXq + (2y � j)H2O/ yCO2 + qX� + kNH3 + (4y � 2j

+ m � 3k)H+ + (4y � 2j + m � 3k � q)e� (1)

Where the elements are represented by their atomic symbols and

X represents a halogen atom, respectively. The stoichiometric

ratio of elements in the organic compound is represented by the

coefficients y, m, j, k and q. The oxidation number (n) in the

complete oxidation process is equal to 4y � 2j + m � 3k � q.

In a typical PeCOD analysis, typical photocurrent responses

of a blank solution containing 2.0 M NaNO3 (iblank, solid line)

and a sample solution containing 2.0 M NaNO3 and organic

compounds were recorded (as shown in Fig. S3†). Accordingly,

Qblank and Qtotal can be obtained by integration of photocur-

rents, iblank and itotal, with time, respectively. i.e.,

Qtotal ¼ðitotaldt (2)

Qblank ¼ðiblankdt (3)

The net charge, Qnet, originated from the oxidation of organic

compounds can be obtained by subtracting Qblank from Qtotal

(see the shaded area in Fig. S3†):

Qnet ¼ Qtotal � Qblank (4)

The net charge (Qnet) can be used to quantify the COD value of

sample according to the following equation:4

COD�mg L�1;O2

� ¼ Qnet

FV� 8000 (5)

where Qnet is the charge of oxidation of organics; F is the

Faraday constant; and V is the volume of solution, a constant for

a given thin-layer reactor.

Fig. 1 (a) SEM image and (b) cross-sectional SEM image of the TiO2/

BDD electrode; (c) XRD patterns of the TiO2/BDD electrode (curve 1),

TiO2/ITO electrode (curve 2) and the pure BDD electrode (curve 3); (d)

J–V plot of the BDD (curve 1) and TiO2/BDD (curve 2) electrodes in the

direct electric measurement.

3. Results and discussion

3.1. Structural characterisation

Anatase and rutile mixed-phase TiO2 electrodes illustrated

higher photocatalytic activities over the oxidation of organic

compounds than pure anatase TiO2 electrodes.33,34 The mixed-

phase TiO2 can be obtained by a calcination process at 700 �C in

air for 2 h. In order to prepare mixed-phase TiO2 on the BDD

substrate, the stability of the BDD electrode was firstly tested at

high temperatures (i.e., 450 �C and 700 �C) in air. The original

BDD has a distinguished crack-free and continuous surface

under SEM (see Fig. S4a†). The BDD surface remained intact

and did not have notable changes in surface morphology and

electrical conductivity after a calcination process at 450 �C in air

for 2 h. However, an incontinuous and corroded BDD surface

was observed under SEM (see Fig S4b†) and the electrical

conductivity was completely lost after the calcination treatment

at 700 �C in air for 0.5 h. These suggest that the BDD could be

damaged by the air at 700 �C. In order to protect the BDD

substrate, Ar atmosphere must be used to replace the oxidative

air. In practice, the dip-coated TiO2/BDD electrodes were cal-

cinated at 450 �C for 0.5 h in air to remove all the carbowax in the

thin film and achieve initial binding among TiO2 particles, and

subsequently sintered at 700 �C for 2 h in the Ar atmosphere.

This journal is ª The Royal Society of Chemistry 2011

Fig. 1a and b show the FESEM images of surface morphology

and cross-section of a typical as-prepared TiO2/BDD electrode,

respectively. Fig. 1a indicates that the TiO2 film was continuous

and uniform. The sizes of the TiO2 nanoparticles were in the

range of 10–20 nm. Fig. 1b suggests that the typical morphology

and nanostructure of BDD (in Fig. S4a†) were well maintained.

In other words, the BDD was not damaged by the above sin-

tering treatments and the aqua regia washing process. The rough

and uneven nanoscale structures are beneficial for the immobi-

lization of the TiO2 film.

XRD patterns were used to analyze the crystalline phase

composition of the TiO2/BDD and TiO2/ITO electrodes. Fig. 1c

shows the XRD patterns of the TiO2/BDD electrode (Curve 1),

the TiO2/ITO electrode (Curve 2) and the pure BDD electrode

(Curve 3) after the 700 �C calcination in Ar. The well-defined

diamond peak at 2q degree of ca. 44� for Curve 1 is almost

identical to that for Curve 3 in Fig. 1c, which further demon-

strates that the physicochemical properties of the BDD were

maintained through the above radical treatments. Another two

distinctive diffraction peaks at 2q degree of ca. 25.4� and 27.6�

for Curve 1 in Fig. 1c could be indexed to anatase-phase (JCPDS

No. 21-1272) and rutile-phase (JCPDS No. 21-1276), respec-

tively. The TiO2/BDD film contained 93% of anatase and 7% of

rutile, while the TiO2/ITO film had a similar ratio (92% of

anatase and 8% of rutile) according to the empirical relationship

used by Depero et al.35 It suggests that the difference of

substrates had no significant effect on the crystalline phase of

TiO2.

Electrical characterization can provide supporting evidence to

demonstrate the formation of the p–n heterojunction.36

Fig. 1d shows electrical current densities, J, recorded upon the

applied potentials after the BDD and the TiO2/BDD electrode

was tested on a micromanipulator manual probe station in the

dark individually. As shown in Fig. 1d, the excellent symmetric

Anal. Methods, 2011, 3, 2003–2009 | 2005

Publ

ishe

d on

01

Aug

ust 2

011.

Dow

nloa

ded

by I

mpe

rial

Col

lege

Lon

don

Lib

rary

on

21/0

5/20

13 0

4:54

:31.

View Article Online

J–V behaviour passing through the zero point indicated that the

contact between the silicon substrate and BDD was simply

ohmic-like contact.17 In strong contrast, a very small current was

observed in the reverse potential region (negative potential

region) while the current increased exponentially in the forward

potential region (positive potential region). In particular, the

reverse current density at �5 V is only 0.03 mA cm�2, while the

forward current density is 0.25 mA cm�2 at +5 V. This asym-

metric characteristic is dramatically different from the symmetric

ohmic-like behaviour of a common resistance mentioned above.

In fact, the asymmetric characteristic of the J–V curve in Fig. 1d

has a rectification ratio of ca. 10 which represents the distinct

diode-like behaviour of a typical p–n heterojunction.17 In this

heterojunction, the forward turn-on and reverse breakdown

voltages were ca. 0.5 and ca.�4 V, respectively. It is worth noting

that the substrate of the BDD is p-type monocrystalline silicon.

Hence, the p–n junction cannot form between the p-type silicon

substrate and p-type BDD. This suggests that the p–n hetero-

junction was established between the TiO2 and the BDD. It is

expected that this can effectively separate the photoelectron/hole

pairs generated under UV light illumination, and enhance the

quantum yield of TiO2 as a photocatalyst.17

3.2. Photoelectrochemical characterization

In order to achieve a maximum photocatalytic efficiency, the

photoelectrons need to be separated from the photoholes.

Application of a positive potential on the semiconductor was

proven to be an effective way to remove the electrons to the

external circuit and achieve this goal.30 However, a too positive

potential may lead to direct electrochemical reactions which will

distort the analytical signals and may even lead to false

measurement. In order to determine the optimal applied poten-

tial during the detection of COD, the effect of applied potential

on the photocurrent signals was first investigated by LSV. Fig. 2

shows the LSVs obtained at the TiO2/BDD electrode and TiO2/

ITO electrode in 0.1 M NaNO3 electrolyte solution containing

10.0 mM of glucose. A high concentration of glucose (10 mM)

was selected as a photohole scavenger to eliminate the influence

of mass transport.30 At both the electrodes, the dark current is

Fig. 2 Linear scanning voltammograms of the TiO2/BDD and TiO2/

ITO electrodes in 0.1 M NaNO3 solution containing 10.0 mM glucose at

a scanning rate of 5 mV s�1.

2006 | Anal. Methods, 2011, 3, 2003–2009

negligible, which suggests water and glucose cannot be directly

decomposed in the applied potential range (�0.5 to +0.65 V).

Under a given UV intensity, the photocurrent increased with the

increased potential before levelling off to a saturated photocur-

rent. Each voltammogram consists of two parts: the increasing

part (i.e. �0.5 to 0 V) and the saturation part (i.e. 0 to +0.65 V).

The increasing part suggests that electron transport in the TiO2/

BDD heterojunction electrodes was the reaction rate-limiting

process, which is similar to the response of a resistor under

varying potentials. The slopes of linear part for the TiO2/BDD

electrode obtained under the experimental conditions are higher

than that of the TiO2/ITO electrode. Therefore, though the

electrochemical resistance of BDD electrode is commonly higher

than that of ITO, the photoelectrochemical resistance of the

TiO2/BDD electrode is lower that that of TiO2/ITO, demon-

strating the high resistance of BDD did not affect the photo-

electrocatalytic activity. The saturated part implies that the

potential applied on the TiO2/BDD electrode was capable of

removing all the electrons generated in the photocatalytic process

at the TiO2 surface.

Typical saturated photocurrents at different electrodes could

be observed when the sweeping potential was greater than 0.0 V,

suggesting that the applied potential should be controlled over

0.0 V to maximize the photoelectrocatalytic efficiency of both

electrodes. Furthermore, the saturation photocurrent of the

TiO2/BDD heterojunction electrode was also higher than that of

TiO2/ITO, suggesting that the photoelectrocatalytic activities of

the TiO2/BDD electrode exceeded that of the TiO2/ITO elec-

trode. Therefore, to ensure a sufficient potential under various

conditions and, at the same time, to minimize the direct elec-

trochemical reaction, a potential of +0.40 V was selected for the

determination of COD.

3.3. Acidoresistance

The acidoresistance of a sensor could be a very important

property for water quality monitoring, especially for the indus-

trial wastewaters containing corrosive substance, such as acidic

pollutants or acidic matrix.37 Sensors based on ITO and Ti metal

substrate have a low acidoresistance due to their chemical nature

of oxide and metal, respectively. This limits their practical

application range and diminishes their reliabilities in online and/

or real time water quality monitoring. In contrast, BDD has an

excellent acid resistance29 evidenced by the fact the distinct BDD

surface was maintained after the aqua regia washing. Therefore,

BDD could be an ideal material to tackle this problem. The

endurability of the TiO2/BDD electrode in the acidic environ-

ment was investigated against the TiO2/ITO electrode.

Herein, both electrodes were polarized at a cyclic sweeping rate

of 5 mV s�1 from �0.5 to +0.65 V in a solution of pH ¼ 1 for 60

cycles. Two series of cyclic voltammograms were obtained and

the saturation photocurrents attained at +0.4 V after various

cycles were used to represent the stability of the electrodes as

shown in Fig. 3. It can be seen that the saturated photocurrent of

TiO2/ITO electrode sharply dropped with the increase of

potential scanning cycles, and reached nearly zero after 40

scanning cycles, indicating the complete loss in conductivity.

This could be visually confirmed by observing the peel-off of

TiO2 film away from the ITO substrate due to the dissolution of

This journal is ª The Royal Society of Chemistry 2011

Fig. 3 Saturation currents obtained from the cyclic voltammograms of

the TiO2/BDD and TiO2/ITO electrodes at +0.4 V in, pH 1, 0.1 M

NaNO3 supporting electrolyte solution after various potential scanning

cycles.

Publ

ishe

d on

01

Aug

ust 2

011.

Dow

nloa

ded

by I

mpe

rial

Col

lege

Lon

don

Lib

rary

on

21/0

5/20

13 0

4:54

:31.

View Article Online

ITO film in the acidic solution. In contrast, the saturated

photocurrents of the TiO2/BDD electrode just slightly decreased

in the first 20 cycles, and kept very steady at 66 mA subsequently,

demonstrating the resilient acid resistance property of the TiO2/

BDD. It can be expected that the TiO2/BDD electrodes would be

more convenient in practical applications and have much

broader applications than the TiO2/ITO electrodes.

3.4. Determination of COD

3.4.1. Validation of analytical principle. Owing to the differ-

ence in chemical identity and structure, the degradation reaction

mechanisms and reaction rates would vary in the photo-

electrocatalytic degradation of different organics.31 This can be

reflected by the different photocurrent profiles in the course of

the degradation of various organic compounds. Fig. 4 shows

a set of photocurrent–time profiles obtained during an exhaus-

tive degradation of 50 mg L�1 organics in the photo-

electrochemical thin-layer cell using the TiO2/BDD electrode at

an applied potential of +0.40 V and under the constant 6.6 mW

cm�2 UV illumination. It can be observed that different organics

possess different photocurrent profiles. Different photocurrent

profiles suggest that the degradation reaction follows different

Fig. 4 Photocurrent profiles of 50 mg L�1 COD of different organics in 2

M NaNO3 solution.

This journal is ª The Royal Society of Chemistry 2011

reaction pathways. In particular, the photocurrent profiles of the

small molecular organic acid (i.e. glycine, glutaric acid, succinic

acid and malonic acid) declined monotonically upon the illumi-

nation until steady-state photocurrents were reached. In strong

contrast, for the aromatic organics (i.e. p-chlorophenol and

KHP), the photocurrents declined rapidly in a short time (ca. 10

s), then unexpectedly increased and finally declined until steady

states were reached, resulting in a shoulder peak-like profile. The

alcoholic compounds (i.e. glucose and methanol) also exhibit

a shoulder peak-like photocurrent profiles but with a wider and

lower ‘‘shoulder’’. The appearance of the shoulder peaks was

mainly due to the generation of intermediates and the polymer-

isation of the intermediates during the degradation process.38

This affects the degradation kinetics evidenced by the fact that

the time required for attainment of the steady state was different.

Interestingly, the Qnet values of the 50 mg L�1 COD of various

organic compounds obtained from Fig. 4 were almost identical,

2020 mC � 4%. This demonstrates that the polymerization

reactions and their effect on the degradation kinetics will not

affect the Qnet values because all the organic compounds will be

degraded eventually according to eqn (1). In other words, the

Qnet obtained from the photocurrent profile of the same COD

value are the same regardless of the identities of the organic

compounds in the wastewater.

In order to further validate the analytical principle, the

PeCOD system incorporated with the TiO2/BDD electrode was

used to detect synthetic samples prepared with a series of organic

compounds with known theoretical COD values. Fig. 5a shows

that the Qnet obtained were directly proportional to the

concentration for all the organic compounds. The oxidation

numbers (n) according to eqn (1) for methanol, glycine, malonic

acid, succinic acid, glutaric acid, glucose, p-chlorophenol, and

KHP are 6, 6, 8, 14, 20, 24, 26, and 30, respectively when these

organics are exhaustively oxidized. The CODTh can be obtained

by multiplying the molar concentrations (CM, mol per L) with its

corresponding oxidation numbers, i.e.,

CODTh (mg L�1) ¼ 8000nCM (6)

Subsequently, the Qnet can be normalized with the CODTh

replacing the CM as shown in Fig. 5b. And the PeCOD can be

acquired by eqn (5), as shown in Fig. 5c. Excellent linear rela-

tionship, a slope of 0.9946 and R2 ¼ 0.9977 between the PeCOD

and CODTh of the organic compounds was obtained, suggesting

that the TiO2/BDD electrode is able to indiscriminately oxidize

various types of organic compounds.

3.4.2. Effect of pH. Real samples containing organic pollut-

ants usually have various pH. pH, however, influences the pho-

tocatalytic oxidation efficiency at the TiO2 electrodes by shifting

the flat band potential and affecting the stability of the TiO2

electrode.27 The effect of pH on the analytical performance was

investigated. KHP was selected as a target analyte because KHP

is a commonly used organic standard for analytical measure-

ments, such as COD and TOC analysis. Fig. 6 shows the effect of

pH on Qnet and Qblank of 50 mg L�1 KHP in the thin-layer cell

using the TiO2/BDD electrode. When the pH was lower than 2,

the Qblank values were slightly lower than that at pH 2–10. The

Qnet were fairly stable from pH 1 to pH 10, indicating that the

Anal. Methods, 2011, 3, 2003–2009 | 2007

Fig. 5 (a) The relationship between the Qnet and the molar concentra-

tion; (b) the relationship between the Qnet and the theoretical COD; (c)

the correlation between the PECOD and theoretical COD.

Fig. 6 Effect of pH on the photocurrent Qnet and Qblank values of the

TiO2/BDD electrode in the degradation of 50 mg L�1 KHP using the

TiO2/BDD electrode.

Fig. 7 The correlation between the PeCOD and the standard COD value

of real samples. 1–3 are the samples collected from surface water; 4–7 are

the samples collected from municipal wastewater plants; 8–12 are the

samples collected from sugar plants; 13–16 are the samples collected from

the food production plants; 17–20 are the samples collected from the

pharmaceutical plants (pH ¼ 1.5–5).

Publ

ishe

d on

01

Aug

ust 2

011.

Dow

nloa

ded

by I

mpe

rial

Col

lege

Lon

don

Lib

rary

on

21/0

5/20

13 0

4:54

:31.

View Article Online

TiO2/BDD electrode is insensitive to the change of solution pH.

For the solution where the pH is higher than 11.0, the Qblank

dramatically increased while the Qnet also changed significantly,

which was due to dramatic increment of water oxidation at high

pH,39 evidenced by the production of oxygen and hydrogen

bubbles. The pH outside the range of pH 1–10 will dramatically

distort the analytical signal and therefore the pH of a sample

needs to be regulated. Overall, the suitable pH range of the TiO2/

BDD electrode lies in the range of pH 1–10, which is wider than

that of the TiO2/ITO (i.e., pH 4–10).39

3.4.3. Analyses of real samples. To investigate the feasibility

of practical application, a TiO2/BDD heterojunction sensor was

2008 | Anal. Methods, 2011, 3, 2003–2009

used to detect numerous real samples obtained from different

sources, including surface water, municipal wastewater and

industrial wastewater (sugar plants, food production plants, and

pharmaceutical plants). Considering the TiO2/BDD electrode

has a very strong acidoresistance and is insensitive to the change

of pH in the range of 1–10, and the pH values of the real samples

were in the range of 1.5 to 9, the pH of these samples were not

adjusted during the PeCOD analysis process. For the purpose of

comparison and analytical method validation, standard COD

values of these samples were also carried out by using the

conventional standard dichromate method.40 The Pearson

correlation method was used to investigate the correlation

between the PeCOD and standard COD values, because it is

commonly considered as a more reliable method where both x

and y axes involve errors of measurement. Fig. 7 shows the

Pearson correlation of the COD values obtained from these two

methods. The COD values of these samples obtained from the

TiO2/BDD electrode were in excellent agreement (R2 ¼ 0.994,

n¼ 20) with that from the conventional dichromate method. The

slope of the principle axis of the correlation ellipse was nearly 1.0.

This journal is ª The Royal Society of Chemistry 2011

Publ

ishe

d on

01

Aug

ust 2

011.

Dow

nloa

ded

by I

mpe

rial

Col

lege

Lon

don

Lib

rary

on

21/0

5/20

13 0

4:54

:31.

View Article Online

This confirms that the pH has an insignificant effect on the COD

value in the PeCOD measurement for these samples using the

TiO2/BDD electrode.

During the real sample analysis, it was observed that the

colour of the samples had insignificant effect on the accuracy of

the PeCOD measurement, this is probably because the change of

UV absorption is trivial due to the use of a thin-layer cell, and the

fact that the variation of UV absorption might affect the pho-

toelectrocatalytic oxidation rate but will not affect the overall

charge for the mineralization reaction.

The practical detection limit of 0.12 mg L�1 COD with a linear

range up to 300 mg L�1 COD can be achieved under the exper-

imental conditions employed using the TiO2/BDD hetero-

junction electrode. The TiO2/BDD electrode is very robust under

normal conditions. Stability experiments suggest that 95–105%

of the net charge for 50 mg L�1 COD glucose was maintained

after the electrode had been used continuously for 500

measurements in 2 months. The analytical results are highly

reproducible with a RSD% value of 1.5% for 11 measurements of

50 mg L�1 COD glucose.

Conclusions

The TiO2/BDD electrodes were prepared by dip-coating the TiO2

nanoparticles onto the BDD substrate and the subsequent

thermal treatments. The PeCOD system with TiO2/BDD elec-

trode has similar analytical performance (in terms of detection

limit, linear range, sensitivity and accuracy and precision) as the

previous works with TiO2/ITO electrodes. As a sensor, the TiO2/

BDD electrode has numerous advantages such as significantly

longer life time, wider application range (i.e., low pH samples),

low background current, strong acidoresistance and improved

photocatalytic activity in comparison with ITO and Ti electrode

substrates. The preliminary results in this work validate appli-

cability of the TiO2/BDD electrode to the PeCOD system using

synthetic and real samples, and demonstrated that the TiO2/

BDD electrode is an ideal sensor for determination of COD

values in wastewater in a rapid, sensitive and accurate fashion.

Acknowledgements

The authors acknowledge the financial support of the ARC

discovery and ARC Future fellowship from Australian Research

Council. And Yanhe Han thanks the financial support from the

Open Funds for Key Laboratory of Industrial Ecology and

Environmental Engineering, China Ministry of Education

(Dalian University of technology).

Notes and references

1 M. R. Hoffmann, S. T. Martin, W. Choi and D. W. Bahnemann,Chem. Rev., 1995, 95, 69–96.

2 N. Watanabe, S. Horikoshi, A. Kawasaki, H. Hidaka andN. Serpone, Environ. Sci. Technol., 2005, 39, 2320–2326.

3 Y. Li, X. Liu, H. Yuan and D. Xiao, Biosens. Bioelectron., 2009, 24,3706–3710.

This journal is ª The Royal Society of Chemistry 2011

4 H. Zhao, D. Jiang, S. Zhang, K. Catterall and R. John, Anal. Chem.,2004, 76, 155–160.

5 Y. Yang, X. Wang and L. Li, J. Am. Ceram. Soc., 2008, 91, 3086–3089.

6 X. Chen, G. Chen, F. Gao and L. Yue Po, Environ. Sci. Technol.,2003, 37, 5021–5026.

7 Y.-h. Cui, X.-y. Li and G. Chen, Water Res., 2009, 43, 1968–1976.8 L. Guo and G. Chen, J. Electrochem. Soc., 2007, 154, D657–D661.9 C. Batchelor-McAuley, C. E. Banks, A. O. Simm, T. G. J. Jones andR. G. Compton, Analyst, 2006, 131, 106–110.

10 K. B. Holt, A. J. Bard, Y. Show and G. M. Swain, J. Phys. Chem. B,2004, 108, 15117–15127.

11 N. Mitani and Y. Einaga, J. Electroanal. Chem., 2009, 626, 156–160.12 W. C. Poh, K. P. Loh, W. D. Zhang, S. Triparthy, J.-S. Ye and

F.-S. Sheu, Langmuir, 2004, 20, 5484–5492.13 I. Sir�es, E. Brillas, G. Cerisola and M. Panizza, J. Electroanal. Chem.,

2008, 613, 151–159.14 R. Kalish, Carbon, 1999, 37, 781–785.15 A. Manivannan, N. Spataru, K. Arihara and A. Fujishima,

Electrochem. Solid-State Lett., 2005, 8, C138–C140.16 J. Qu and X. Zhao, Environ. Sci. Technol., 2008, 42, 4934–4939.17 H. Yu, S. Chen, X. Quan, H. Zhao and Y. Zhang, Environ. Sci.

Technol., 2008, 42, 3791–3796.18 A. P. H. Association, A. W. W. Association and W. E. Federation,

Apha-Awwa-Wef, Washington, D.C., 19th edn, 1995.19 K.-H. Lee, T. Ishikawa, S. J. McNiven, Y. Nomura, A. Hiratsuka,

S. Sasaki, Y. Arikawa and I. Karube, Anal. Chim. Acta, 1999, 398,161–171.

20 C. Silva, C. D. C. Conceicao, V. Bonifacio, O. Fatibello andM. Teixeira, J. Solid State Electrochem., 2009, 13, 665–669.

21 S. Ai, M. Gao, Y. Yang, J. Li and L. Jin, Electroanalysis, 2004, 16,404–409.

22 J. Li, L. Li, L. Zheng, Y. Xian, S. Ai and L. Jin, Anal. Chim. Acta,2005, 548, 199–204.

23 J. Q. Li, L. P. Li, L. Zheng, Y. Z. Xian and L. T. Jin, Meas. Sci.Technol., 2006, 17, 1995–2000.

24 H. Yu, H. Wang, X. Quan, S. Chen and Y. Zhang, Electrochem.Commun., 2007, 9, 2280–2285.

25 P. Westbroek and E. Temmerman, Anal. Chim. Acta, 2001, 437, 95–105.

26 S. Zhang, D. Jiang and H. Zhao, Environ. Sci. Technol., 2006, 40,2363–2368.

27 S. Zhang, H. Zhao, D. Jiang and R. John, Anal. Chim. Acta, 2004,514, 89–97.

28 J. Zhang, B. Zhou, Q. Zheng, J. Li, J. Bai, Y. Liu and W. Cai, WaterRes., 2009, 43, 1986–1992.

29 J. Stotter, Y. Show, S. Wang and G. Swain, Chem. Mater., 2005, 17,4880–4888.

30 D. Jiang, H. Zhao, S. Zhang and R. John, J. Phys. Chem. B, 2003,107, 12774–12780.

31 Y. Han, S. Zhang, H. Zhao, W. Wen, H. Zhang, H. Wang andF. Peng, Langmuir, 2010, 26, 6033–6040.

32 W. Wen, H. Zhao and S. Zhang, J. Phys. Chem. C, 2009, 113, 10830–10832.

33 D. Jiang, S. Zhang andH. Zhao,Environ. Sci. Technol., 2007, 41, 303–308.

34 Y. Zhang, J. Chen and X. Li, Catal. Lett., 2010, 139, 129–133.35 L. E. Depero, L. Sangaletti, B. Allieri, E. Bontempi, R. Salari,

M. Zocchi, C. Casale and M. Notaro, J. Mater. Res., 1998, 13,1644–1649.

36 J. Yuan, H. Li, S. Gao, Y. Lin and H. Li, Chem. Commun., 2010, 46,3119–3121.

37 G. Capar, L. Yilmaz andU. Yetis, J.Membr. Sci., 2006, 281, 560–569.38 D. S. Muggli, J. T. McCue and J. L. Falconer, J. Catal., 1998, 173,

470–483.39 S. Zhang, L. Li, H. Zhao and G. Li, Sens. Actuators, B, 2009, 141,

634–640.40 Q. Zheng, B. Zhou, J. Bai, L. Li, Z. Jin, J. Zhang, J. Li, Y. Liu, W. Cai

and X. Zhu, Adv. Mater. (Weinheim, Ger.), 2008, 20, 1044–1049.

Anal. Methods, 2011, 3, 2003–2009 | 2009