understanding anodic wear at boron doped diamond film ... · understanding anodic wear at boron...

10
Electrochimica Acta 89 (2013) 122–131 Contents lists available at SciVerse ScienceDirect Electrochimica Acta jou rn al h om epa ge: www.elsevier.com/locate/electacta Understanding anodic wear at boron doped diamond film electrodes Brian P. Chaplin a,, David K. Hubler b , James Farrell b a Department of Civil and Environmental Engineering and Villanova Center for the Advancement of Sustainable Engineering Villanova University, Villanova, PA 19085, United States b Department of Chemical and Environmental Engineering University of Arizona, Tucson, AZ 85721, United States a r t i c l e i n f o Article history: Received 14 September 2012 Received in revised form 28 October 2012 Accepted 28 October 2012 Available online 23 November 2012 Keywords: Boron-doped diamond Anodic corrosion Density functional theory Ultrananocrystalline a b s t r a c t This research investigated the mechanisms associated with anodic wear of boron-doped diamond (BDD) film electrodes. Cyclic voltammetry (CV), x-ray photoelectron spectroscopy (XPS), scanning electron microscopy (SEM), and electrochemical impedance spectroscopy (EIS) were used to measure changes in electrode response and surface chemistry as a function of the charge passed and applied current den- sity. Density functional theory (DFT) modeling was used to evaluate possible reaction mechanisms. The initial hydrogen-terminated surface was electrochemically oxidized at lower potentials than water oxi- dation (1.83 V/SHE), and was not catalyzed by the hydrogen-terminated surface. In the region where water oxidation produces hydroxyl radicals (OH ), the hydrogen-terminated surface may also be oxi- dized by chemical reaction with OH . Oxygen atoms became incorporated into the surface via reaction of carbon atoms with OH , forming both C O and C-OH functional groups, that were also detected by XPS measurements. Experimental and DFT modeling results indicate that the oxygenated diamond surface lowers the potential for activationless water oxidation from 2.74 V/SHE for the hydrogen terminated sur- face to 2.29 V/SHE for the oxygenated surface. Electrode wear was accelerated at high current densities (i.e., 500 mA cm 2 ), where SEM results indicated oxidation of the BDD film resulted in significant surface roughening. These results are supported by EIS measurements that document an increase in the double- layer capacitance as a function of the charge passed. DFT simulations provide a possible mechanism that explains the observed diamond oxidation. DFT simulation results indicate that BDD edge sites (=CH 2 ) can be converted to COOH functional groups, which are further oxidized via reactions with OH to form H 2 CO 3(aq.) with an activation energy of 58.9 kJ mol 1 . © 2012 Elsevier Ltd. All rights reserved. 1. Introduction Boron-doped diamond (BDD) film electrodes have been exten- sively studied for their ability to oxidize a wide range of water contaminants. Contaminants can be oxidized by a combination of direct electron transfer and reaction with hydroxyl radicals (OH ) produced from water oxidation [1–5]. BDD electrodes have also been reported to possess high stability under anodic polarization [6–9], which has motivated the use of these electrodes in various emerging water treatment and sensor applications [10–12]. The traditional metric used to test electrode stability is the accel- erated lifetime test (ALT). The ALT employs very high anodic current densities (e.g., 1 A cm 2 ), and monitors the time until electrode failure. Results from the ALT have shown that film delamination is the primary failure mode of BDD electrodes, and thus extensive research has focused on increasing the adhesion of BDD films to var- ious substrates [8,13–16]. As a result of this work, reported lifetimes for BDD electrodes as high as 804 Ah cm 2 have been achieved, as Corresponding author. Tel.: +1 610 519 4967; fax: +1 610 519 4967. E-mail address: [email protected] (B.P. Chaplin). measured by the ALT [15]. Although the ALT is useful for compar- ing the relative stability of various electrodes under high anodic current densities, the conditions of the ALT are not representative of water treatment applications that employ much lower current densities (e.g., < 20 mA cm 2 ). Therefore, the ALT could force failure modes that do not occur under conditions used in water treatment applications. Although BDD electrodes are generally considered to have a high anodic stability, several studies indicate that measurable wear of the BDD film can occur upon high, applied anodic charges (e.g., >100 Ah cm 2 ) and current densities (e.g., 1—2 A cm 2 ) [16–19]. Experimental evidence suggests that BDD resistance to wear is enhanced with increasing boron-doping levels [17], and wear is accelerated in the presence of organic electrolytes (e.g., acetic acid) [18] and alkaline conditions [20,21]. However, the specific mech- anisms associated with anodic wear on BDD electrodes have not been thoroughly explored. Various studies have shown that anodic polarization under low current densities (e.g., < 20 mA cm 2 ) results in oxygen incor- poration into the BDD surface, which may represent a starting point for BDD oxidation. X-ray photoelectron spectroscopy (XPS) measurements indicate that freshly prepared BDD electrodes are 0013-4686/$ see front matter © 2012 Elsevier Ltd. All rights reserved. http://dx.doi.org/10.1016/j.electacta.2012.10.166

Upload: nguyentruc

Post on 25-Apr-2018

232 views

Category:

Documents


5 download

TRANSCRIPT

Page 1: Understanding anodic wear at boron doped diamond film ... · Understanding anodic wear at boron doped ... These results are supported by EIS measurements that document an increase

U

Ba

b

a

ARRAA

KBADU

1

scdpb[e

edfirif

0h

Electrochimica Acta 89 (2013) 122– 131

Contents lists available at SciVerse ScienceDirect

Electrochimica Acta

jou rn al h om epa ge: www.elsev ier .com/ locate /e lec tac ta

nderstanding anodic wear at boron doped diamond film electrodes

rian P. Chaplina,∗, David K. Hublerb, James Farrellb

Department of Civil and Environmental Engineering and Villanova Center for the Advancement of Sustainable Engineering Villanova University, Villanova, PA 19085, United StatesDepartment of Chemical and Environmental Engineering University of Arizona, Tucson, AZ 85721, United States

r t i c l e i n f o

rticle history:eceived 14 September 2012eceived in revised form 28 October 2012ccepted 28 October 2012vailable online 23 November 2012

eywords:oron-doped diamondnodic corrosionensity functional theoryltrananocrystalline

a b s t r a c t

This research investigated the mechanisms associated with anodic wear of boron-doped diamond (BDD)film electrodes. Cyclic voltammetry (CV), x-ray photoelectron spectroscopy (XPS), scanning electronmicroscopy (SEM), and electrochemical impedance spectroscopy (EIS) were used to measure changesin electrode response and surface chemistry as a function of the charge passed and applied current den-sity. Density functional theory (DFT) modeling was used to evaluate possible reaction mechanisms. Theinitial hydrogen-terminated surface was electrochemically oxidized at lower potentials than water oxi-dation (≤ 1.83 V/SHE), and was not catalyzed by the hydrogen-terminated surface. In the region wherewater oxidation produces hydroxyl radicals (OH•), the hydrogen-terminated surface may also be oxi-dized by chemical reaction with OH•. Oxygen atoms became incorporated into the surface via reaction ofcarbon atoms with OH•, forming both C O and C-OH functional groups, that were also detected by XPSmeasurements. Experimental and DFT modeling results indicate that the oxygenated diamond surfacelowers the potential for activationless water oxidation from 2.74 V/SHE for the hydrogen terminated sur-face to 2.29 V/SHE for the oxygenated surface. Electrode wear was accelerated at high current densities

−2

(i.e., 500 mA cm ), where SEM results indicated oxidation of the BDD film resulted in significant surfaceroughening. These results are supported by EIS measurements that document an increase in the double-layer capacitance as a function of the charge passed. DFT simulations provide a possible mechanism thatexplains the observed diamond oxidation. DFT simulation results indicate that BDD edge sites (=CH2)can be converted to COOH functional groups, which are further oxidized via reactions with OH• to formH2CO3(aq.) with an activation energy of 58.9 kJ mol−1.

. Introduction

Boron-doped diamond (BDD) film electrodes have been exten-ively studied for their ability to oxidize a wide range of waterontaminants. Contaminants can be oxidized by a combination ofirect electron transfer and reaction with hydroxyl radicals (OH•)roduced from water oxidation [1–5]. BDD electrodes have alsoeen reported to possess high stability under anodic polarization6–9], which has motivated the use of these electrodes in variousmerging water treatment and sensor applications [10–12].

The traditional metric used to test electrode stability is the accel-rated lifetime test (ALT). The ALT employs very high anodic currentensities (e.g., ≥ 1 A cm−2), and monitors the time until electrodeailure. Results from the ALT have shown that film delaminations the primary failure mode of BDD electrodes, and thus extensive

esearch has focused on increasing the adhesion of BDD films to var-ous substrates [8,13–16]. As a result of this work, reported lifetimesor BDD electrodes as high as 804 Ah cm−2 have been achieved, as

∗ Corresponding author. Tel.: +1 610 519 4967; fax: +1 610 519 4967.E-mail address: [email protected] (B.P. Chaplin).

013-4686/$ – see front matter © 2012 Elsevier Ltd. All rights reserved.ttp://dx.doi.org/10.1016/j.electacta.2012.10.166

© 2012 Elsevier Ltd. All rights reserved.

measured by the ALT [15]. Although the ALT is useful for compar-ing the relative stability of various electrodes under high anodiccurrent densities, the conditions of the ALT are not representativeof water treatment applications that employ much lower currentdensities (e.g., < 20 mA cm−2). Therefore, the ALT could force failuremodes that do not occur under conditions used in water treatmentapplications.

Although BDD electrodes are generally considered to have a highanodic stability, several studies indicate that measurable wear ofthe BDD film can occur upon high, applied anodic charges (e.g.,>100 Ah cm−2) and current densities (e.g., 1—2 A cm−2) [16–19].Experimental evidence suggests that BDD resistance to wear isenhanced with increasing boron-doping levels [17], and wear isaccelerated in the presence of organic electrolytes (e.g., acetic acid)[18] and alkaline conditions [20,21]. However, the specific mech-anisms associated with anodic wear on BDD electrodes have notbeen thoroughly explored.

Various studies have shown that anodic polarization under

low current densities (e.g., < 20 mA cm−2) results in oxygen incor-poration into the BDD surface, which may represent a startingpoint for BDD oxidation. X-ray photoelectron spectroscopy (XPS)measurements indicate that freshly prepared BDD electrodes are
Page 2: Understanding anodic wear at boron doped diamond film ... · Understanding anodic wear at boron doped ... These results are supported by EIS measurements that document an increase

himic

hactstsotwisTe

tatpcttwlcov(bnctod

2

2

1Rbt1tm

2

oBea

2

Piwcwte

B.P. Chaplin et al. / Electroc

ydrogen terminated [22–25], and these hydrogen terminationsre easily converted to oxygen-containing functional groups (e.g.,arboxyl, carbonyl, and hydroxyl groups) upon anodic polariza-ion [23–28]. The functional groups on the BDD surface have atrong effect on charge transfer in both anodic and cathodic reac-ions [25,29–32]. The formation of oxygenated groups has beenhown to both inhibit some reactions while facilitating other typesf reactions [25,26,29–32]. The electrochemical response of solu-ion phase redox couples has also been reported to depend onhether the electrode surface was anodically or cathodically polar-

zed prior to the measurement [16,19,33,34], suggesting that theurface functional groups continuously evolve on the BDD surface.his changing surface chemistry can affect the reproducibility oflectrode performance in water treatment and sensor applications.

Despite evidence from ALTs indicating that BDD film delamina-ion is the primary failure mode at high anodic current densitiesnd charges, at low anodic current densities studies have shownhat the BDD surface becomes oxidized, suggesting BDD wear isossible under conditions encountered in water treatment appli-ations. Since BDD oxidation can affect the performance of waterreatment and sensor applications, a fundamental understanding ofhe mechanisms of BDD wear is needed. Thus, the aim of this workas to characterize the wear mechanisms of BDD electrodes during

ong-term anodic polarization. Specifically, changes in the electro-hemical response of BDD electrodes were assessed as a functionf the cumulative charge passed using a combination of cyclicoltammetry (CV) and electrochemical impedance spectroscopyEIS) measurements. Changes in surface termination were assessedy XPS, and surface morphology changes were assessed by scan-ing electron microscopy (SEM). Density functional theory (DFT)alculations were used to interpret the experimental results, ando gain insight into important mechanisms related to the formationf oxygenated-functional groups and the wear of BDD electrodesuring anodic polarization.

. Materials and Methods

.1. Electrodes

All electrodes consisted of ultrananocrystalline BDD films on.0 cm2 p-silicon substrates (Advanced Diamond Technologies,omeoville, IL). The BDD films were deposited on the substratesy chemical vapor deposition (CVD). The BDD films were growno a thickness of approximately 2 �m under conditions of 750-2000 parts-per-million of trimethyl borane in flowing CH4 atemperatures between 700—800◦C. Prior to use, the BDD films had

easured resistivities between 0.05—0.1 ohm-cm.

.2. Reagents

All chemicals were used as received (Sigma-Aldrich) with-ut any additional purification and were of reagent grade purity.ackground electrolyte solutions of 1 M NaClO4 were used for allxperiments, as this electrolyte has been shown to be nonreactivet BDD anodes and cathodes [30,35].

.3. Electrode Ageing Experiments

Electrodes were aged under anodic polarization in 200 mLyrex® beakers. Circular 1 cm2 BDD disk electrodes were mountednto custom PEEK® electrode holders. Electrical contact was made

ith the backside of each electrode using a 316 stainless steel

urrent collector contained within the electrode holder, and aater-tight seal was accomplished by sealing the face of the elec-

rode with a Viton® gasket. The cell holder provided a 0.35 cm2 areaxposed to the solution. The BDD electrode was used as the anode

a Acta 89 (2013) 122– 131 123

and was held stationary in the center of the beaker. A 10 cm by18 cm 316 stainless steel 18 mesh screen (TWP Inc., Berkeley, CA)was used as the cathode and was wrapped around the inner circum-ference of the beaker. The background electrolyte was 1 M NaClO4and the solution was continuously stirred using a magnetic Teflon-coated stir bar. The electrodes were connected to a MASTECH®(Hong Kong) model HY1803D galvanostatic power supply operatedwithout a reference electrode, and polarized at constant anodic cur-rent densities of 0, 5, 50 and 500 mA cm−2, termed from here on asBDD-0, BDD-5, BDD-50, and BDD-500, respectively.

2.4. Electrochemical Characterization Methods

BDD films were characterized electrochemically using CV andEIS. The temperature was controlled at 22 ◦C using a circulatingwater bath. Currents and electrode potentials were controlled andmonitored using a Gamry Series G 750 potentiostat/galvanostat.The BDD electrodes were mounted in the electrode holdersdescribed previously. A Princeton Applied Research (PAR) model316 rotating disk electrode (RDE) assembly was used to rotate theelectrodes at 3000 rpm during EIS measurements, and the elec-trode was held stationary during CV scans. The counter electrodewas a 12 cm long by 0.3 mm diameter Pt wire and the referenceelectrode was a PAR Hg/Hg2SO4 electrode saturated with K2SO4.Potentials are reported versus the standard hydrogen electrode(SHE). To prevent contamination of the electrode surface fromadsorbed organic compounds from the atmosphere, the BDD elec-trode was cleaned via anodic polarization in the electrolyte solutionat a current density of 20 mA cm−2 for 10 minutes before eachset of measurements. CV experiments were conducted in a 1 MNaClO4 background electrolyte in the absence and presence of5 mM K3Fe(CN)6 and 5 mM K4Fe(CN)6, and the potential was sweptat a scan rate of 100 mV s−1.

EIS experiments were conducted in a 1 M NaClO4 backgroundelectrolyte containing 5 mM K3Fe(CN)6 and 5 mM K4Fe(CN)6. EISmeasurements were made at the open circuit potential (∼0.465 V/SHE) with an amplitude of ±10 mV and over a frequencyrange of 0.1 to 3 × 105 Hz. EIS data was fit by equivalent circuitmodels using Echem AnalystTMV 5.6.1a software (Gamry Instru-ments). Average values for double-layer capacitances (Cdl (F cm−2))were extracted from values determined for constant phase ele-ments according to methods by Brug et al. [36]. The total impedance(Z) for the equivalent circuit model can be expressed as:

Z = Rs + 1(jω�T

) (1)

where RS is the solution resistance (ohm cm2); j is the notationfor the imaginary unit; � is the radial frequency (radians s−1); �is related to the angle of rotation of a purely capacitive line onthe complex plane plots (dimensionless), and is equal to 0 for apure resistor and 1 for a pure capacitor; and T (F cm−2 s�-1) can beexpressed according to equation (2) [36].

T = C�dl

[R−1S + R−1

ct ]1−�

(2)

Rct represents the resistance to charge transfer (ohm cm2) for aFaradaic reaction.

2.5. Quantum Mechanical Simulations

Density functional theory (DFT) simulations were performed togain insights into the reaction behavior of the functional groups on

the BDD surface and to investigate the mechanisms associated withelectrode oxidation. Two structures were used to model the BDDelectrode surface: a previously-described 10-carbon atom clusterterminated by hydrogen atoms [30], and a previously-described
Page 3: Understanding anodic wear at boron doped diamond film ... · Understanding anodic wear at boron doped ... These results are supported by EIS measurements that document an increase

1 himica Acta 89 (2013) 122– 131

ha[swBcoiiu

tAftgtoestutduspm

2

zXtswD

3

3

emacsicoisfao

TtFFBw

Fig. 1. CV scans of the BDD-0 electrode: a) First CV scan 0.64—3.0 V/SHE, b) Fifth

24 B.P. Chaplin et al. / Electroc

ydrogen terminated 34-carbon structure with a central borontom [37]. All DFT calculations were performed with the DMol337–39] package in the Accelrys Materials Studio [40] modelinguite on a personal computer. All simulations used double-numericith polarization (DNP) basis sets [41] and the gradient-correctedecke-Lee-Yang-Parr (BLYP) [42,43] functionals for exchange andorrelation. The nuclei and core electrons were described by DFTptimized semilocal pseudopotentials [44]. Implicit solvation wasncorporated into all simulations by use of the COSMO-ibs polar-zable continuum model [45]. Thermal smearing of 0.005 Ha wassed for numerical convergence.

The activation energies (Ea) for direct electron transfer as a func-ion of the electrode potential were calculated by the method ofnderson and Kang [46]. Reactant energies were calculated as a

unction of the reaction coordinate, defined as the bond length ofhe bond that was formed or broken by the reaction. Product ener-ies were calculated using the atomic positions determined fromhe optimized reactant structures, followed by self-consistent fieldptimization of the electronic configurations. The energy of the freelectron on the vacuum scale was adjusted to the SHE scale byubtracting 4.6 eV [46]. Product energies were adjusted to reflecthe electrode potential by shifting the energy profile of the prod-ct species downwards by 96.5 kJ mol−1 (i.e., 1.0 eV) to increasehe electrode potential by 1.0 V and upwards by 96.5 kJ mol−1 toecrease the electrode potential by 1.0 V. Intersection of the prod-ct and reactant energy profiles yields the bond length of transitiontate and the Ea for the reaction. Simulations involving solvatedrotons included 2 explicit water molecules to more accuratelyimic physical reality [47].

.6. Analytical Methods

XPS measurements were performed at the University of Ari-ona Laboratory for Electron Spectroscopy and Surface Analysis.PS measurements were not conducted as a function of time that

he BDD electrode was exposed to air after anodic ageing, so anyurficial chemistry changes that may have occurred during this timeere not documented. SEM images were performed at Advancediamond Technologies (Romeoville, IL).

. Results and Discussion

.1. Oxidation of C-H Surface Sites

Initial CV scans in 1 M NaClO4 are shown in Fig. 1 for the BDD-0lectrode that was subjected to the anodic surface cleaning treat-ent only (20 mA cm−2 for 10 minutes). The first anodic scan shows

distinct, broad oxidation peak at 2.38 V (Fig. 1a), which indi-ates that a significant number of reduced sites (e.g., C-H) weretill present on the BDD surface after the anodic surface clean-ng treatment. However, the peak at 2.38 V disappeared after fiveonsecutive anodic scans, indicating that the reduced sites werexidized (Fig. 1b). This peak reappeared when the CV scans werenitially swept in the cathodic direction (Fig. 1c). After three con-ecutive scans (-2 V to 3 V), the size of the peak at 2.38 V showed nourther change (Fig. 1d), and an additional smaller peak at 1.44 Vppeared (Fig. 1e). The cathodic scans also showed the appearancef a reduction peak at -1.4 V.

DFT modeling was used to interpret the CV scans shown in Fig. 1.wo clusters based on the structure of diamond were used in ordero investigate mechanisms of BDD surface oxidation, as shown in

ig. 2. The structure in Fig. 2a includes a B atom and the structure inig. 2b does not. These structures include three distinct CHx sites:

C-H and C-H sites present on the BDD facets, and CH2 siteshich would likely be present at diamond grain boundaries. These

CV scan 0.64—3.0 V/SHE, c) First CV scan 0.64 to -1.84 to 3.0 V/SHE, d) Third CV scan0.64 to -1.84 to 3.0 V/SHE. e) Blowup of CV scan d). Dashed Line shows oxidationpeak at 2.38 V/SHE.

sites exist on freshly-prepared BDD electrodes, and are expectedto be electroactive [48]. DFT simulations were used to investigatethe following reactions that may occur on a freshly-prepared BDDelectrode:

C-H → C• + H+ + e- (3)

CH2→ C•H + H+ + e- (4)

B C-H → B C• + H+ + e- (5)

The activation barriers for reactions (3)-(5) as a function ofelectrode potential are shown in Fig. 3. The activation barriersfor reactions (3), (4), and (5) become comparable to the thermalenergy (RT = 2.45 kJ mol−1) at potentials of 1.83, 1.79, and 0.74 V,respectively, indicating that these reactions could readily occur atand above these potentials. Near these potentials, mA level currentbegan to flow, as shown in Fig. 1a. The disappearance of the broadpeak in Fig. 1b can likely be attributed to depletion of the oxidiz-able H atoms on the electrode surface. The absence of a measurablepeak corresponding to reaction (5) may be attributed to the lowconcentration of these sites on the electrode surface.

The DFT modeling also indicates that loss of H atoms fromthe electrode surface resulted in the formation of double bondsbetween adjacent C atoms (discussed in detail in Section 3.5). Thedata in Fig. 1c and d indicate that surface oxidation resulting in Hatom loss was at least partially reversible upon cathodic polariza-tion. The production of atomic hydrogen by the Volmer dischargereaction (H+ + e− → H•) produces atomic hydrogen that can reactwith the surface C atoms. The large separation between the anodic

and cathodic peaks in Fig. 1d indicates that the oxidation andreduction reactions involved a multi-step mechanism where bondbreakage and formation occurred. The exact mechanism of this pro-cess has not been previously determined, and is beyond the scope
Page 4: Understanding anodic wear at boron doped diamond film ... · Understanding anodic wear at boron doped ... These results are supported by EIS measurements that document an increase

B.P. Chaplin et al. / Electrochimica Acta 89 (2013) 122– 131 125

F ted dH ersionr

otpfstocfs

Dt

H

Tpbpbgwte

F

ig. 2. Structures used to simulate the BDD surface. a) Hydrogen (white)-terminaydrogen-terminated diamond structure containing 10 carbon atoms. See online v

eader is referred to the web version of the article.).

f the current study. However, previous studies have attributedhis behavior to the oxidation/reduction of sp2 C basal and edgelanes [49], an increase in carrier concentration due to subsur-ace hydrogen incorporation during cathodic treatment [29], andurface roughening due to electrochemical oxidation [29]. Addi-ional studies using XPS measurements indicate that a portion ofxygenated functional groups (i.e., C OH) can be removed uponathodic polarization resulting in a corresponding increase in CHx

unctional groups [25]. The above processes are not mutually exclu-ive, and all may contribute to the behavior shown in Fig. 1.

Hydroxyl radical attack of CHx sites was also investigated usingFT simulations. Hydroxyl radicals are produced via water oxida-

ion according to:

2O → OH• + H+ + e- (6)

he activation barrier for this reaction as a function of electrodeotential is shown in Fig. 4. At potentials ≥ 2.74 V, the activationarrier became comparable to RT, and thus the reaction can readilyroceed at room temperature. The modeling results are confirmedy the data in Fig. 1 where the currents for O2 evolution becomereater than 1 mA at potentials > 2.7 V. The fact that significant

ater oxidation did not take place at potentials < 2.7 V indicates

hat oxidation of surface functional groups by OH• occurs only afterlectrochemical oxidation of CHx sites via reactions (3)-(5). Once

ig. 3. Activation barriers for reactions 3-5 as a function of electrode potential.

iamond structure containing 34 carbon atoms (gray) and 1 boron atom (pink). b) for colors. (For interpretation of the references to color in this figure legend, the

OH• are produced, the DFT modeling indicates that the removal ofan H atom from the BDD surface via reactions (7)–(9):

C-H + OH• → C• + H2O (7)

CH2 + OH• → C•H + H2O (8)

B C-H + OH• → B C• + H2O (9)

proceeds without an activation barrier. Energy is released ineach reaction, with �E = -112.1 kJ mol−1 for reaction (7), �E = -103.2 kJ mol−1 for reaction (8), and �E = -265.7 kJ mol−1 forreaction (9).

3.2. Formation of Oxygenated Functional Groups

The surface carbon radicals formed in reactions (3)-(5) and (7)-(9) are not stable, and thus may react with water or OH• to formoxygenated functional groups. Experimental evidence indicates theelectrochemical response of the BDD electrodes changed dramat-ically as a result of extended anodic polarization. CV scans for theBDD-5 electrode are shown in Fig. 5 after passing a charge of 1 Ahcm−2. Three distinct oxidation peaks at 1.44, 1.88, and 2.49 V wereobserved prior to oxygen evolution. After a charge of 10 Ah cm−2

was passed, only the peak at 1.88 V remained.In order to obtain a measure of the functional groups present on

the surface of the BDD electrode as a result of anodic oxidation, XPSanalysis was performed on electrodes BDD-5 (10 Ah cm−2), BDD-50

Fig. 4. Activation energy for uncatalyzed (solid line) and catalyzed (dashed line)water oxidation as a function of electrode potential.

Page 5: Understanding anodic wear at boron doped diamond film ... · Understanding anodic wear at boron doped ... These results are supported by EIS measurements that document an increase

126 B.P. Chaplin et al. / Electrochimic

-100

0

100

-4.0 0.0 4.0

a)

b)

Fig. 5. CV scan of the BDD-5 electrode at applied charges of a) 1.0 Ah cm−2, b) 10 Ahcm−2. Scan rate of 100 mV s−1 in 1 M NaClO4. CV was initially scanned in cathodicdirection. Inset shows entire potential range of CV scan.

Fig. 6. Surface functional groups determined by XPS spectra. Carbon percentagesdc

(lFisf1stctg

atttfo

B

oxidation of =CH2 to form C O (as shown in reactions (4), (11),

etermined at applied charges of 0.01 (BDD-0), 10 (BDD-5), 50 (BDD-50), and 100 Ahm−2 (BDD-500).

50 Ah cm−2), BDD-500 (100 Ah cm−2), and BDD-0, which was ana-yzed by XPS prior to the electrode-cleaning step. Results shown inig. 6 indicate that the control electrode (BDD-0), consisted primar-ly of sp3 carbon (C-C and CHx), with minor amounts of graphiticp2 carbon (5.2%) and a significant fraction of oxygenated diamondunctional groups. For example, C-OH groups were estimated at5.5%, and C O groups were estimated at 2.2%. These oxygenatedites presumably form via exposure of the electrode to oxygen inhe atmosphere [34]. Upon anodic polarization to different appliedharges, the primary changes relative to BDD-0 were a decrease inhe CHx and C-C sp3 content, and a corresponding increase in oxy-enated functional groups ( C-OH and C O), as shown in Fig. 6.

DFT simulations were also used to investigate the mechanismsssociated with the formation of C O and C-OH surface func-ional groups that were detected by XPS. DFT results indicate thathe C•, C•H, and B≡C• sites, generated as discussed above (reac-ions 7-9), will react without an activation barrier with OH• toorm C-OH, CH-OH, and B C-OH sites at potentials where waterxidation occurs, according to the following reactions:

C• + OH• → C-OH (10)

C•H + OH• → CH-OH (11)

C• + OH• → B C-OH (12)

a Acta 89 (2013) 122– 131

These reactions release energy, with �E = -276.4 kJ mol−1 forreaction (10), �E = -294.9 kJ mol−1 for reaction (11), and �E = -118.1 kJ mol−1 for reaction (12).

The formation of C O groups on the BDD surface was inves-tigated at =CH-OH sites on the BDD cluster which was formed viareaction (11). Formation of C O groups requires the removal oftwo H atoms, which is simulated electrochemically in the followingdirect electron transfer reactions, in which the hydrogen bonded tothe carbon is removed in the first step:

CH-OH → C•-OH + H+ + e- (13)

C•-OH → C O + H+ + e- (14)

The order of hydrogen removal may be reversed so that the hydro-gen bonded to the oxygen is removed in the first step:

CH-OH → CH-O• + H+ + e- (15)

CH-O• → C O + H+ + e- (16)

Simulations found that reaction (13) becomes activationless at≥ 1.72 V and reaction (15) becomes activationless at ≥ 1.73 V. Thefinal steps to form C O surface groups are activationless at lowerpotentials than the initial steps, and should thus readily occur afterreactions (13) or (15). Reaction (14) becomes activationless at ≥-0.60 V and reaction (16) becomes activationless at ≥ 1.55 V.

Because reactions (13) and (15) become activationless at poten-tials similar to reactions (3) and (4), the individual reactions wouldnot be distinguishable in the CV scans. However, Figs. 1 and 5 showan oxidation peak at 1.44 V, which may be due to the formationof C O sites according to reaction (16). The peak current densityfor the freshly prepared electrode was 0.3 mA cm−2 (Fig. 1e), andwas 5 mA cm−2 for the BDD-5 electrode after passing a 1.0 Ah cm−2

charge (Fig. 5a). The growth of the peak with anodic ageing supportsthe proposed mechanism for C O formation. The peak at 1.44 Vwas not present after an applied charge of 10 Ah cm−2 (Fig. 5b),indicating that much of the available =CH2 sites had already beenconverted to C O at this electrode age. The peak does not reap-pear with cathodic polarization, indicating the conversion to C Ois not reversible at the cathodic polarizations used. Results alsoindicate that the formation of C O sites begins to stabilize thesurface with respect to further electrochemical reactions. The sta-bility of the C O site to oxidation reactions is investigated in moredetail in Section 3.6.

Once the BDD surface becomes oxidized additional direct elec-tron transfer reactions can occur that produce radical sites on theBDD surface. Reactions at C-OH and B C-OH sites are shownbelow:

C-OH → C-O• + H+ + e- (17)

B C-OH → B C-O• + H+ + e- (18)

Simulations found that reaction (17) becomes activationless atpotentials ≥ 1.74 V and reaction (18) becomes activationless at ≥0.80 V. These reactions may also contribute to the oxidation peaksshown in Figs. 1 and 5. The oxygen radical sites have also beensuggested to be important active sites for compound oxidation onBDD electrodes [31].

The peaks at 1.44 V and 1.88—2.49 V shown in Figs. 1 and 2 havebeen observed in previous studies of BDD electrodes, and wereattributed to oxidation of non-diamond C impurities present at thegrain boundaries of the electrode, originating from the CVD process[16,19,49–51]. The =CH2 sites used for DFT modeling are repre-sentative sites that are present in non-diamond C, and thus the

and (13-16)) are consistent with the oxidation of non-diamond C aswell as diamond edge sites. Other oxidation reactions, such as theoxidation of hydroquinone-like structures on the surface, have also

Page 6: Understanding anodic wear at boron doped diamond film ... · Understanding anodic wear at boron doped ... These results are supported by EIS measurements that document an increase

himica Acta 89 (2013) 122– 131 127

bdtCootoodpa

3

treB

wrudtrtFd

wsfapamws(

3

ltwdf∼cwscApcTit

aaTp

a)

b)

c)

-5

0

5

10

0.8 1.3 1.8 2.3 2.8

a)

b)c)

Fig. 7. CV scan of electrode BDD-50 at applied charges of a) 1 Ah cm−2, b) 10 Ah cm−2

and c) 50 Ah cm−2. Scan rate of 100 mV s−1 in 1 M NaClO4. CV was initially scannedin cathodic direction. Inset shows blowup of a-c.

20 mA cm-2

a)

b)

c)

Electrode poten�a l/V SHE

B.P. Chaplin et al. / Electroc

een proposed [19,51]. DFT simulations indicate hydroquinone oxi-ation becomes activationless at potentials ≥ 0.72 V, which is closeo the potential at which �A levels of current started flowing in theV scans in Figs. 1 and 5, indicating that these sites are not abundantn the BDD surface. Oxidation of hydroquinone-like sites and otherxygenated sites, but not fully oxidized, sites (e.g., =CH-O• (reac-ion 16)) could also explain the peak at 1.44 V, because this peaknly appears on the aged electrodes. Therefore, the oxidation peaksbserved in Figs. 1 and 5a may be attributed to oxidation of bothiamond and non-diamond C impurities. The reactions above alsorovide a feasible pathway that explains the formation of C-OHnd C O sites on BDD electrodes during anodic polarization.

.3. Effect of Surface Termination on Water Oxidation

Results shown in the inset of Fig. 5 clearly show that increasinghe applied charge from 1.0 to 10 Ah cm−2 for the BDD-5 electrodeesulted in a significant reduction in the overpotential for oxygenvolution. This suggests that oxygen atoms incorporated into theDD surface may catalyze the water oxidation reaction.

DFT simulations were used to investigate the mechanisms ofater oxidation on the BDD surface, according to reaction (6). The

eaction is initially examined independent of the BDD surface, sim-lating a hydrophobic surface (e.g., C-H, B C-H, and =CH2) thatoes not interact with the water molecules. Fig. 4 shows the Ea forhe uncatalyzed reaction as a function of electrode potential. Theeaction is activationless at potentials ≥ 2.74 V, close to the poten-ial at which current rapidly increased for a fresh electrode (seeig. 1). This result suggests that for a fresh electrode, the surfaceoes not catalyze water oxidation.

To investigate how a more oxygenated surface may catalyzeater oxidation, H2C O and H3CO• molecules were added to the

ystem to simulate surface C O, B C-O•, and C-O• sites. The Ea

or the H2C O catalyzed water oxidation reaction is shown in Fig. 4s a function of electrode potential, and becomes activationless atotentials ≥ 2.29 V. The lower activation barrier for H2C O cat-lyzed water oxidation indicates that oxygenated functional groupsay catalyze the water oxidation reaction. A similar catalytic effectas found for H3CO•. Surface site catalysis is consistent with the

hift of water oxidation to lower potentials that can be seen in Fig. 5inset) after electrode aging.

.4. Electrode Wear Under High Current Density

Prior work has shown that very high anodic polarization canead to observable wear of BDD electrodes [16–19]. To inves-igate the mechanisms of electrode oxidation, BDD electrodesere subjected to current densities of 50 and 500 mA cm−2 forifferent applied charges. CV scans for the BDD-50 electrode per-ormed in 1 M NaClO4 show that the peak originally present at

2.38 V was greatly attenuated after an applied charge of 1 Ahm−2 was passed (Fig. 7a). After an applied charge of 10 Ah cm−2

as passed, an increased separation of the forward and reversecans was observed. This separation corresponds to increasedapacitance/pseudo-capacitance of the electrode (Fig. 7b and inset).fter a charge of 50 Ah cm−2 was passed, an oxidation/reductioneak couple appeared at ∼ 1.9 V, and a significant increase inapacitance/pseudo-capacitance was once again observed (Fig. 7c).he overpotential for oxygen evolution also decreased withncreased applied charge, which is consistent with the experimen-al and DFT modeling results discussed in Section 3.3.

CV scans for BDD-500 in 1 M NaClO4 are shown in Fig. 8. After

n applied charge of 1 Ah cm−2, a current response was observedt ∼ 1.5 V that extended until the end of the anodic scan at 3.0 V.his electrochemical response is different than that observed forrevious electrodes tested (i.e., BDD-5 and BDD-50). The very broad

Fig. 8. CV scans of BDD-500 electrode at applied charges of a) 1.0 Ah cm−2, b) 50 Ahcm−2, and c) 100 Ah cm−2. Scan rate of 100 mV s−1 in 1 M NaClO4. CV was initiallyscanned in cathodic direction.

current response may be attributed to a severe distortion of the dia-mond lattice, which likely occurred due to the high, applied currentdensity of 500 mA cm−2. The oxidation of C-C bonds in the diamondlattice may have caused bond breakage and the creation of defectsites, which would both increase the surface area of the electrodeand create electroactive sites with varying redox potentials. How-ever, at higher applied charges (i.e., 50 and 100 Ah cm−2) the peakat potentials >2.0 V was no longer observed (Fig. 8b-c). Instead, atan applied charge of 50 Ah cm−2 a peak was observed at ∼ 1.9 V inthe CV scan, and at an applied charge of 100 Ah cm−2 the CV scanshowed an increased separation between the forward and reversescans, indicative of an electrode with high capacitance/pseudo-capacitance [52]. The lack of distinct peaks in the CV scan afterpassing an applied charge of 100 Ah cm−2, suggests that the elec-troactive sites on the BDD surface were fully oxidized. That is, thesp2 C content was oxidized to CO2, and the diamond sites wereconverted to oxygen functionalities (e.g., C O and C-OH).

SEM analysis was conducted to determine if physical changesoccurred to the BDD electrodes as a result of anodic polarization.Significant changes in surface morphologies were observed forBDD-500. The SEM micrographs shown in Fig. 9 compare the sur-face morphologies of BDD-500 at locations exposed to 0 and 100 Ahcm−2 applied charges. It is apparent from the results that the elec-trode subjected to anodic polarization (100 Ah cm−2) underwentphysical changes as a result of the electrochemical treatment. The

black areas in the SEM micrograph represent areas on the BDD elec-trode where the diamond has been etched away, and the whiteareas represent areas of raised topography. These results have beenobserved in prior studies with BDD electrodes subjected to current
Page 7: Understanding anodic wear at boron doped diamond film ... · Understanding anodic wear at boron doped ... These results are supported by EIS measurements that document an increase

128 B.P. Chaplin et al. / Electrochimica Acta 89 (2013) 122– 131

de at

dXtariwghssaowrcs

3

sgsstmcmoAt

ra

ta

Fig. 9. Scanning electron micrographs of BDD-500 electro

ensities of 1.0 A cm−2 [16–19]. The SEM results help to explainPS results for the BDD-500 electrode in Fig. 6. The BDD-500 elec-

rode did not show a significant decrease in C-C content after anodicgeing, which may be attributed to an increase in surface areaelated to the morphology changes observed in the SEM micrographn Fig. 9. The diamond crystal size of the freshly prepared BDD films

as on the order of approximately 2-5 nm, and etching of the BDDrain boundaries and edge sites (=CH2) during anodic ageing mayave decreased the grain size and exposed more underlying C-Cites. Since the depth of penetration of XPS is on the nanometercale [53], the BDD-500 electrode may contain smaller crystal sizend therefore XPS data may reflect more of the bulk compositionf the BDD film relative to the freshly prepared electrode. Otherork has shown that low anodic current densities (100 �A cm−2)

esulted in more oxygenated functional groups than higher anodicurrent densities (100 mA cm−2) [25], which is similar to our resultshown in Fig. 6.

.5. Mechanism for sp2 Carbon Formation

DFT modeling results indicate that sp2 C is formed on the BDDurface when two adjacent carbon atoms lose their attached hydro-en atoms and reconstruct to minimize the dangling bonds. For thetructures shown in Fig. 2, this can occur at the =CH2 and B C-Hites (each with a neighboring C-H site). It was found that the ini-ial step in sp2 C formation is the removal of a hydrogen atom, which

ay be accomplished electrochemically, as in reactions (3)-(5), orhemically, as in reactions (7)-(9). In both cases, the B C-H site isost reactive, followed by the C-H site. The energy for removal

f a second hydrogen atom may be affected by removal of the first.djacent to a B C• site, the possible reactions for the removal of

he second hydrogen atom are shown in reactions (19) and (20).

C-H → C• + H+ + e- (19)

C-H + OH• → C• + H2O (20)

Reaction (19) becomes activationless at potentials ≥ 0.43 V, andeaction (20) is activationless with �E = -236 kJ mol−1. Adjacent to

C• site, the removal reaction for the second hydrogen is:

• •

CH2 + OH → C H + H2O (21)

Reaction (21) is activationless with �E = -53.6 kJ mol−1. In reac-ions (19) and (20), the second oxidation step occurs more easilydjacent to a C• site than from a fully hydrogen-terminated

0 Ah cm−2 (left) and 100 Ah cm−2 (right) applied charges.

surface (=C-H and =CH2), since the surface reconstructs to forma stable double bond. This mechanism may be reversible and thuscould help to explain the CV results shown in Fig. 1, but detailed DFTcalculations regarding this mechanism are beyond the scope of thisstudy.

3.6. Mechanism for BDD Electrode Wear

DFT simulations were also performed to understand mecha-nisms responsible for electrode wear, as observed in SEM imagespresented in Fig. 9. The C O site was chosen for further study, asit was detected on oxidized electrodes by XPS (Fig. 6), and previouswork with graphite anodes suggests this site can be further oxidizedto a–COOH site [54,55]. DFT simulations indicate that the oxidationof a C O site by OH• forms a =CO•OH adduct, by addition of OH•

to the C atom, as shown in the following reaction.

C O + OH• → CO•OH (22)

Reaction (22) has Ea = 27.1 kJ mol−1 and �E = -69.5 kJ mol−1. Therelatively low Ea and the negative reaction energy indicate that thisreaction is feasible at room temperature, and likely occurred duringour experiments.

Further reactions were explored at the =CO•OH site, which pro-vided a feasible mechanism for etching of the diamond surface(Fig. 10). As shown in Fig. 10, the =CO•OH site (labeled as C1) recon-structs to form a –COOH site, which is further oxidized to formaqueous H2CO3. The process of reconstruction to a –COOH site firstrequires the oxidative removal of a hydrogen atom from an adjacentcarbon atom, creating a C• site (labeled as C2). Hydrogen removalcan occur either electrochemically, as in reactions (3) and (5), orchemically, as in reactions (7) and (9). The energy for removal of thehydrogen atom is affected by the functional groups of the neigh-boring C atoms. DFT simulations indicate that hydrogen removalby the electrochemical pathway was activationless at potentialsbetween 0.43 and 1.83 V (depending on neighboring C atom func-tionality), and the chemical pathway was activationless under allconditions tested. Once the hydrogen atom is removed from the C2carbon, the surface reconstructs to minimize the dangling bonds.

The two carbons atoms (labeled as C2 and C3) previously bondedto the =CO•OH carbon, form a bond to each other and a C-COOHsite is formed. The C2 and C3 carbon atoms remain in their origi-nal sp3 hybridization. In the final step, an OH• attack at the –COOH
Page 8: Understanding anodic wear at boron doped diamond film ... · Understanding anodic wear at boron doped ... These results are supported by EIS measurements that document an increase

B.P. Chaplin et al. / Electrochimica Acta 89 (2013) 122– 131 129

F . The ia Key: ct .).

sd

�cao(tet5ipdct(gos

owwaetHep

3

tt

TS

ig. 10. Proposed mechanism for oxidation of diamond surface at a carbon edge sitet carbon atoms C2 and C3. Refer to the text for details of the reaction mechanism.

o color in this figure legend, the reader is referred to the web version of the article

ite results in the formation of H2CO3, which dissociates from theiamond surface into solution, as shown in reaction (23).

COOH + OH• → H2CO3(aq.) (23)

Reaction (23) has Ea = 58.9 kJ mol−1 and a reaction energy,E = −64.3 kJ/mol. The relatively high Ea of 58.9 kJ mol−1, indi-

ates that this process will proceed slowly at room temperature,nd explains why electrode wear of BDD electrodes was notbserved under ambient temperatures and low current densities5—50 mA cm−2) in this study. The above mechanism also explainshe results from previous studies that document BDD oxidation orlectrochemical polishing at high anodic potentials and elevatedemperatures (40—60 ◦C) [17–19]. Based on the calculated Ea of8.9 kJ mol−1, increasing the temperature from 20 ◦C to 60 ◦C will

ncrease the rate of reaction (23) by a factor of 18. Although the tem-erature of our experiments was controlled at 20 ◦C, ohmic heatinguring ageing of the BDD-500 electrode may have occurred, whichaused an increase in the temperature of the electrode relativeo the electrolyte solution and thus increased the rate of reaction23). XPS results did not detect the formation of–COOH functionalroups, but results shown in Fig. 6 show a decrease in the contentf C O functional groups with increasing current density, thusupporting the mechanism of conversion of C O to H2CO3(aq.).

Several studies have shown that electrolysis in the presence ofrganics [18] or alkaline conditions [20,21] greatly enhances BDDear rates. Although the effects of solution conditions on electrodeear is beyond the scope of our study, the formation of H2CO3(aq.) as

product in reaction (23) suggests that alkaline conditions wouldnhance electrode wear. The H2CO3(aq.) formed would be convertedo CO3

2− at elevated pH, and push reaction (23) towards products.owever, a more detailed study is necessary to support this hypoth-sis and explore the mechanism of enhanced electrode wear in theresence of organics.

.7. EIS Analysis

EIS experiments were also conducted to more fully characterizehe electrochemical response of the BDD electrodes as a func-ion of anodic ageing by measuring values for Rct and Cdl. EIS was

able 1ummary of equivalent circuit model fit to EIS data. Model is presented in Eqs. (1) and (2

Electrode Applied Charge (Ah cm −2) Rs (ohm cm2) Rc

BDD-0 0.01 9.1 10BDD-5 1.0 11 44BDD-5 10 12 44BDD-50 1.0 19 10BDD-50 10 8.8 37BDD-50 50 11 23BDD-500 10 16 54BDD-500 50 10 22BDD-500 100 13 18

nitial structure on the left contains a =CO•OH site at carbon atom C1, and C-H sitesarbon (gray); hydrogen (white); oxygen (red). (For interpretation of the references

conducted in a 1 M NaClO4 electrolyte containing 5 mM K4Fe(CN)6and 5 mM K3Fe(CN)6. Results of the EIS model fits are shown inTable 1. Several trends are observed from the parameter valuesdetermined from the EIS model fits. The Cdl value for the freshelectrode (BDD-0) was 14.4 �F cm−2. For the electrodes polarizedat 5, 50 and 500 mA cm−2, the Cdl values increased as a functionof the applied charge passed. For applied charges of 10 and 50 Ahcm−2, the increases in Cdl were inversely proportional to the currentdensity used to pass the charge. This indicates that the corrosionmechanism may be more related to the time of operation than tothe applied current density.

The values for the Cdl began to plateau between applied chargesof 50 and 100 Ah cm−2. A value for Cdl of 3.23 mF cm−2 was found forBDD-500 at 100 Ah cm−2. The large increase in the values of Cdl asa function of applied charge may be related to the electrochem-ical etching of the diamond grain boundaries and non-diamondC, which increased the surface area and the density of chargedsites at the electrode surface [29]. Comparing the Cdl value for theBDD-500 at 100 Ah cm−2 (3.23 mF cm−2) to the Cdl value for theBDD-0 electrode (14.4 �F cm−2), provides an estimate of the elec-troactive surface area for the BDD-500 electrode at 100 Ah cm−2

of 78.5 cm2, which is 224 times greater than the geometric surfacearea of 0.35 cm2. Other studies have found a 15—21 fold increasein the Cdl value upon roughening the microcrystalline BDD sur-face through a variety of methods (e.g., catalytic roughening, steamactivation) [56,57]. The EIS data compliments the CV scans thatshowed high capacitance, and SEM images that showed a rough-ened surface upon anodic ageing. The Cdl value found for BDD-500after an applied charge of 100 Ah cm−2 (3.23 mF cm−2, normalizedby geometric surface area) was much higher than Cdl values forroughened microcrystalline BDD electrodes, where values as highas 165—234 �F cm−2 have been reported [56–58]. Nanocrystallinediamond films grown on micrometer porous silicon resulted in Cdlvalues between 230—990 �F cm−2 [59]. Therefore, the high Cdl val-ues measured in our study may be a result of the nm crystal size

and warrants further investigation.

The Rct values also show interesting trends. The fresh elec-trode (BDD-0) had an Rct value of 10.7 � cm2. Upon anodic ageing,the Rct value initially increased after an applied charge of 1 Ah

).

t (ohm cm2) T (F cm −2 s�−1) � Cdl (�F cm−2)

.7 2.19E−04 0.61 14.37

.2 2.16E−05 0.86 18.43

.4 1.03E−03 0.88 17631 9.39E−05 0.92 168.0.2 6.40E−04 0.89 1043.8 1.59E−03 0.89 2914.4 4.12E−04 0.90 754.3.9 1.76E−03 0.84 2594.4 2.00E−03 0.85 3229

Page 9: Understanding anodic wear at boron doped diamond film ... · Understanding anodic wear at boron doped ... These results are supported by EIS measurements that document an increase

1 himic

ci(tBokvaoaewF

aadaaSgb

ub0cpa

4

eepeorAtiadqo

A

eFI

R

[

[

[

[

[

[

[

[

[

[

[

[

[

[

[

[

[

[

[

[

[

[

30 B.P. Chaplin et al. / Electroc

m−2 was passed on all electrodes, followed by a slow decreasen the Rct value with progressively higher applied anodic charges10—100 Ah cm−2). These trends are likely related to changes inhe surface termination, sp2 C content, and surface area of theDD electrodes. Freshly prepared electrodes have a high contentf hydrogen-terminated sites, which exhibit fast electron transferinetics for the FeCN6

4-/3− redox couple [29]. The increase in the Rct

alue at an applied charge of 1 Ah cm−2 is likely related to oxygen-tion of the BDD surface, which slows the electron transfer kineticsf the FeCN6

4-/3− redox couple [29]. The increase in Rct after anpplied charge of 1 Ah cm−2 was passed could also be related tolectrochemical oxidation of a portion of the sp2 C to CO2. Previousork has shown that sp2 C is active for the electron transfer of the

eCN64-/3− redox couple [49].

Upon further anodic ageing, the Rct values steadily decreased,nd reached a minimum value for the BDD-500 electrode after anpplied charge of 100 Ah cm−2 was passed (18.4 ohm cm2). Theecrease in the Rct value is likely related to an increase in surfacerea and an increase in charged surface sites. An increase in surfacerea can be inferred from the high Cdl value measured by EIS andEM images (Fig. 9), and an increase in charged surface sites is sug-ested by both CV scans (Fig. 8) and the high Cdl value measuredy EIS.

The exponential � in Eqs. (1) and (2) was found to be less thannity, which is attributed to a distribution in time constants causedy electrode surface roughness [60]. Values for � ranged between.84 and 0.92, and in general decreased as a function of the appliedharge passed. The decreasing trend of � with the applied chargeassed indicates that the electrodes became rougher during anodicgeing, which supports CV, SEM, and measured Cdl values.

. Conclusions

The results of this study provide experimental and modelingvidence that explains the formation of oxygenated sites on BDDlectrodes and anodic wear of BDD that has been documented inrior studies [16–19]. It is proposed that freshly prepared BDDlectrodes that contain a high density of CHx surface sites can bexidized by a sequence of direct electron transfer reactions andeactions with OH• to form C O and -C-OH functional groups.t high applied charges (e.g., 100 Ah cm−2) SEM images indicate

hat the BDD electrode experienced significant wear. DFT model-ng results identified possible mechanisms that occur at potentialst and below that of water oxidation that explain the oxidation ofiamond edge sites (=CH2) to H2CO3(aq.). These reactions provideuantitative information that helps to explain the observed wearf BDD electrodes.

cknowledgements

Funding for this work was provided by the National Sci-nce Foundation (CBET-0931749) and the National Scienceoundation Small Business for Innovative Research program (NSF-IP-0945935).

eferences

[1] J.F. Zhi, H.B. Wang, T. Nakashima, T.N. Rao, A. Fujishima, Electrochemical incin-eration of organic pollutants on boron-doped diamond electrode. Evidence fordirect electrochemical oxidation pathway, Journal of Physical Chemistry B 107(2003) 13389.

[2] K.E. Carter, J. Farrell, Oxidative destruction of perfluorooctane sulfonate using

boron-doped diamond film electrodes, Environmental Science and Technology42 (2008) 6111.

[3] A. Kapalka, G. Foti, C. Comninellis, Investigation of the anodic oxidation ofacetic acid on boron-doped diamond electrodes, Journal of the ElectrochemicalSociety 155 (2008) E27.

[

a Acta 89 (2013) 122– 131

[4] B.P. Chaplin, G. Schrader, J. Farrell, Electrochemical oxidation of N-Nitrosodimethylamine with boron-doped diamond film electrodes, Environ-mental Science and Technology 43 (2009) 8302.

[5] B.P. Chaplin, G. Schrader, J. Farrell, Electrochemical destruction of N-Nitrosodimethylamine in reverse osmosis concentrates using boron-dopeddiamond film electrodes, Environmental Science and Technology 44 (2010)4264.

[6] G.M. Swain, The Susceptibility to surface corrosion in acidc fluoride media - Acomparison of diamond, HOPG, and glassy carbon, Journal of the Electrochem-ical Society 141 (1994).

[7] Q.Y. Chen, M.C. Granger, T.E. Lister, G.M. Swain, Morphological and micro-structural stability of boron-doped diamond thin film electrodes in an acidicchloride medium at high anodic current densities, Journal of the Electrochem-ical Society 144 (1997) 3806.

[8] X.M. Chen, G.H. Chen, F.R. Gao, P.L. Yue, High-performance Ti/BDD electrodesfor pollutant oxidation, Environmental Science and Technology 37 (2003) 5021.

[9] X. Chen, F. Gao, G. Chen, Comparison of Ti/BDD and Ti/SnO2–Sb2O5 electrodesfor pollutant oxidation, Journal of Applied Electrochemistry 35 (2005) 185.

10] A. Kraft, Doped diamond: a compact review on a new, versatile electrode mate-rial, International Journal of Electrochemical Science 2 (2007) 355.

11] C. Comninellis, G. Chen, Electrochemistry for the Environment, Springer, NewYork, 2010.

12] O. Chailapakul, W. Siangproh, D.A. Tryk, Boron-doped diamond-based sensors:a review, Sensor Letters 4 (2006) 99.

13] X.M. Chen, F.R. Gao, G.H. Chen, Comparison of Ti/BDD and Ti/SnO2-Sb2O5 elec-trodes for pollutant oxidation, Journal of Applied Electrochemistry 35 (2005)185.

14] Y. Tian, X.M. Chen, C. Shang, G.H. Chen, Active and stable Ti/Si/BDD anodes forelectro-oxidation, Journal of the Electrochemical Society 153 (2006) J80.

15] L. Guo, G. Chen, Long-term stable Ti/BDD electrode fabricated with HFCVDmethod using two-stage substrate temperature, Journal of the ElectrochemicalSociety 154 (2007) D657.

16] B.P. Chaplin, I. Wylie, H. Zheng, J.A. Carlisle, J. Farrell, Characterization of theperformance and failure mechanisms of boron-doped diamond ultrananocrys-talline diamond electrodes, Journal of Applied Electrochemistry 41 (2011)1329.

17] N. Katsuki, E. Takahashi, M. Toyoda, T. Kurosu, M. Iida, S. Wakita, Y. Nishiki, T.Shimamune, Water electrolysis using diamond thin-film electrodes, Journal ofthe Electrochemical Society 145 (1998).

18] M. Panizza, G. Sine, I. Duo, L. Ouattara, C. Comninellis, Electrochemical polishingof boron-doped diamond in organic media, Electrochemical and Solid-StateLetters 6 (2003) D17.

19] I. Duo, C. Levy-Clement, A. Fujishima, C. Comninellis, Electron transfer kinet-ics on boron-doped diamond Part I: influence of anodic treatment, Journal ofApplied Electrochemistry 34 (2004) 935.

20] R. DeClements, G.M. Swain, The formation and electrochemical activity ofmicroporous diamond thin film electrodes in concentrated KOH, Journal ofApplied Electrochemistry 144 (1997) 856.

21] U. Griesbach, D. Zollinger, H. Putter, C. Comninellis, Evaluation of boron dopeddiamond electrodes for organic electrosynthesis on a preparative scale, Journalof Applied Electrochemistry 35 (2005) 1265.

22] H. Martin, A. Argoitia, U. Landau, A.B. Anderson, J.C. Angus, Hydrogen andoxygen evolution on boron-doped diamond electrodes, Journal of the Electro-chemical Society 143 (1996) L133.

23] J. Xu, M.C. Granger, Q. Chen, J.W. Strojek, T.E. Lister, G.M. Swain,Boron-doped diamond thin-film electrodes, Analytical Chemistry 69 (1997)591A.

24] C.H. Goeting, F. Marken, A. Gutierrez-Sosa, R.G. Compton, J.S. Foord, Electro-chemically induced surface modifications of boron-doped diamond electrodes:an X-ray photoelectron spectroscopy study, Diamond and Related Materials 9(2000) 390.

25] H. Girard, N. Simon, D. Ballutaud, M. Herlem, A. Etcheberry, Effect of anodic andcathodic treatments on the charge transfer of boron doped diamond electrodes,Diamond and Related Materials 16 (2007) 316.

26] H. Notsu, I. Yagi, T. Tatsuma, D.A. Tryk, A. Fujishima, Introduction of oxygen-containing functional groups onto diamond electrode surfaces by oxygenplasma and anodic polarization, Electrochemical and Solid-State Letters 2(1999) 522.

27] H. Notsu, I. Yagi, T. Tatsuma, D.A. Tryk, A. Fujishima, Surface carbonyl groupson oxidized diamond electrodes, Journal of Electroanalytical Chemistry 492(2000) 31.

28] H. Notsu, T. Fukazawa, T. Tatsuma, D.A. Tryk, A. Fujishima, Hydroxyl groups onboron-doped diamond electrodes and their modification with a silane couplingagent, Electrochemical and Solid-State Letters 4 (2001) H1.

29] M.C. Granger, G.M. Swain, The influence of surface interactions on thereversibility of ferri/ferrocyanide at boron-doped diamond thin-film elec-trodes, Journal of the Electrochemical Society 146 (1999) 4551.

30] D. Mishra, Z.H. Liao, J. Farrell, Understanding reductive dechlorination oftrichloroethene on boron-doped diamond film electrodes, Environmental Sci-ence and Technology 42 (2008) 9344.

31] O. Azizi, D. Hubler, G. Schrader, J. Farrell, B.P. Chaplin, Mechanism of perchlo-

rate formation on boron-doped diamond anodes, Environmental Science andTechnology 45 (2011) 10582.

32] I. Yagi, H. Notsu, T. Kondo, D.A. Tryk, A. Fujishima, Electrochemical selectiv-ity for redox systems at oxygen-terminated diamond electrodes, Journal ofElectroanalytical Chemistry 473 (1999) 173.

Page 10: Understanding anodic wear at boron doped diamond film ... · Understanding anodic wear at boron doped ... These results are supported by EIS measurements that document an increase

himic

[

[

[

[

[

[

[

[[

[

[

[

[

[

[

[

[

[

[

[

[

[

[

[

[

[

B.P. Chaplin et al. / Electroc

33] H.B. Suffredini, V.A. Pedrosa, L. Codognoto, S.A.S. Machado, R.C. Rocha, L.A.Avaca, Enhanced electrochemical response of boron-doped diamond elec-trodes brought on by a cathodic surface pre-treatment, Electrochimica Acta49 (2004) 4021.

34] G.R. Salazar-Banda, L.S. Andrade, P.A.P. Nascente, P.S. Pizani, R.C. Rocha, L.A.Avaca, On the changing electrochemical behaviour of boron-doped diamondsurfaces with time after cathodic pre-treatments, Electrochimica Acta 51(2006) 4612.

35] J. Iniesta, P.A. Michaud, M. Panizza, G. Cerisola, A. Aldaz, C. Comninellis,Electrochemical oxidation of phenol at boron-doped diamond film electrode,Electrochimica Acta 46 (2001) 3573.

36] G.J. Brug, A.L.G. Vandeneeden, M. Sluytersrehbach, J.H. Sluyters, Theanalysis of electrode impedances complicated by the presense of a con-stant phase element, Journal of Electroanalytical Chemistry 176 (1984)275.

37] Y. Cai, A.B. Anderson, J.C. Angus, L.N. Kostadinov, Hydrogen evolution ondiamond electrodes by the Volmer-Heyrovsky mechanism, Journal of the Elec-trochemical Society 154 (2007) F36.

38] B. Delley, An all-electron numerical-method for solving the local density func-tional for polyatomic-molecules, Journal of Chemical Physics 92 (1990) 508.

39] B. Delley, From molecules to solids with the DMol3 approach, Journal of Chem-ical Physics 113 (2000) 7756.

40] Materials Studio, v.4.2; Accelrys Corporation, San Diego, CA.41] B. Delley, Fast calculation of electrostatics in crystals and large molecules, Jour-

nal of Physical Chemistry 100 (1996) 6107.42] A.D. Becke, Density-functional exchange-energy approximation with correct

asymptotic-behavior, Physical Review A 38 (1988) 3098.43] C.T. Lee, W.T. Yang, R.G. Parr, Development of the Colle-Salvetti correlation-

energy formula into a functional of the electron-density, Physical Review B 37(1988) 785.

44] B. Delley, Hardness conserving semilocal pseudopotentials, Physical Review B66 (2002).

45] B. Delley, The conductor-like screening model for polymers and surfaces,

Molecular Simulation 32 (2006) 117.

46] A.B. Anderson, D.B. Kang, Quantum chemical approach to redox reac-tions including potential dependence: application to a model for hydrogenevolution from diamond, Journal of Physical Chemistry A 102 (1998)5993.

[

[

a Acta 89 (2013) 122– 131 131

47] C.P. Kelly, C.J. Cramer, D.G. Truhlar, Adding explicit solvent molecules to con-tinuum solvent calculations for the calculation of aqueous acid dissociationconstants, Journal of Physical Chemistry A 110 (2006) 2493.

48] H.V. Patten, S.C.S. Lai, J.V. Macpherson, P.R. Unwin, Active sites for outer-sphere, inner-sphere, and complex multistage electrochemical reactions atpolycrystalline boron-doped diamond electrodes (pBDD) revealed with scan-ning electrochemical cell microscopy (SECCM), Analytical Chemistry 84 (2012)5427.

49] E. Mahe, D. Devilliers, C. Comninellis, Electrochemical reactivity at graphiticmicro-domains on polycrystalline boron doped diamond thin-films electrodes,Electrochimica Acta 50 (2005) 2263.

50] M.C. Granger, M. Witek, J. Xu, J. Wang, M. Hupert, A. Hanks, M.D. Koppang,J.E. Butler, G. Lucazeau, M. Mermoux, J.W. Strojek, G.M. Swain, Standard elec-trochemical behavior of high-quality, boron-doped polycrystalline diamondthin-film electrodes, Analytical Chemistry 72 (2000) 3793.

51] A. Kapalka, G. Foti, C. Comninellis, Investigations of electrochemical oxygentransfer reaction on boron-doped diamond electrodes, Electrochimica Acta 53(2007) 1954.

52] A.J. Bard, L.R. Faulkner, Electrochemical Methods: Fundamentals and Applica-tions, John Wiley and Sons, 1980.

53] M. Hosokawa, K. Nogi, M. Naito, T. Yokohama, Nanoparticle Technology Hand-book, Elsevier, Amsterdam, 2012.

54] C.U. Pittman, W. Jiang, Z.R. Yue, S. Gardner, L. Wang, H. Toghiani, C. Leon, Surfaceproperties of electrochemically oxidized carbon fibers, Carbon 37 (1999).

55] M. Rueffer, D. Bejan, N.J. Bunce, Graphite: an active or an inactive anode? Elec-trochimica Acta 56 (2011) 2246.

56] T. Ohashi, W. Sugimoto, Y. Takasu, Catalytic roughening of surface layers of BDDfor various applications, Electrochimica Acta 54 (2009) 5223.

57] T. Ohashi, J.F. Zhang, Y. Takasu, W. Sugimoto, Steam activation of boron dopeddiamond electrodes, Electrochimica Acta 56 (2011) 5599.

58] T. Watanabe, T.K. Shimizu, Y. Tateyama, Y. Kim, M. Kawai, Y. Einaga, Giantelectric double-layer capacitance of heavily boron-doped diamond electrode,Diamond and Related Materials 19 (2010) 772.

59] N.G. Ferreira, A.F. Azevedo, A.F. Beloto, M. Amaral, F.A. Almeida, F.J. Oliveira,R.F. Silva, Nanodiamond films growth on porous silicon substrates for electro-chemical applications, Diamond and Related Materials 14 (2005) 441.

60] T. Pajkossy, Impedance of rough capacitive electrodes, Journal of Electroana-lytical Chemistry 364 (1994) 111.