Vous êtes sur la page 1sur 13

Forum Article

www.acsami.org

Structure, Electronic Properties, and Electrochemical Behavior of a


Boron-Doped Diamond/Quartz Optically Transparent Electrode
Naihara Wac hter, Catherine Munson, Romana Jarosova, Isil Berkun, Timothy Hogan,
Romeu C. Rocha-Filho, and Greg M. Swain*,

Departamento de Qumica, Universidade Federal de Sao Carlos, C. P. 676, 13560-970 Sao Carlos, SP, Brazil
Department of Chemistry, Michigan State University, East Lansing, Michigan 48824, United States

Department of Computer and Electrical Engineering, Michigan State University, East Lansing, Michigan 48824, United States

ABSTRACT: The morphology, microstructure, chemistry, electronic properties, and electrochemical behavior of a boron-doped nanocrystalline diamond (BDD) thin lm grown on quartz
were evaluated. Diamond optically transparent electrodes (OTEs) are useful for transmission
spectroelectrochemical measurements, oering excellent stability during anodic and cathodic
polarization and exposure to a variety of chemical environments. We report on the
characterization of a BDD OTE by atomic force microscopy, optical spectroscopy, Raman
spectroscopic mapping, alternating-current Hall eect measurements, X-ray photoelectron
spectroscopy, and electrochemical methods. The results reported herein provide the rst
comprehensive study of the relationship between the physical and chemical structure and
electronic properties of a diamond OTE and the electrodes electrochemical activity.
KEYWORDS: optically transparent diamond electrode, optical, chemical, and electronic properties, electrochemical characterization,
ionic liquids

1. INTRODUCTION
The use of diamond as an optically transparent electrode
(OTE) for spectroelectrochemical measurements was rst
reported back in the early 2000s.1,2 Our group and others
have made use of either free-standing diamond plates or thin
lms of boron-doped diamond (BDD) deposited on quartz for
transmission spectroelectrochemical measurements in the UV
vis region of the electromagnetic spectrum1,36 or on undoped
Si for measurements in the mid- to far-IR region.2,7 In our
opinion, the use of optically transparent diamond electrodes in
spectroelectrochemistry is an underdeveloped area of research
with this material. BDD possesses attractive qualities as an
OTE: a wide optical window, a wide working potential window
(>3 V in aqueous media), low background current, microstructural stability during anodic and cathodic polarization, and
resistance to molecular adsorption and fouling.1,48 Depending
on the doping level and lm thickness, diamond lms are
transparent in the visible (300700 nm) and mid- to far-IR
(<1100 cm1) regions. The preparation of an OTE requires
balances of lm thickness (thick enough to be continuous but
not too thick to signicantly reduce transmission) and doping
level (sucient to impart electrical conductivity but not so high
as to make the lm opaque). In the case of diamond, one has
some control over the materials optical properties through
selection of the chemical vapor deposition (CVD) conditions
used for thin-lm growth. The wide optical window coupled
with the interesting electrochemical properties make BDD
OTEs unique materials for transmission spectroelectrochemistry and sensor technologies in which both electrochemical and
spectroscopic signal transduction is desired.
XXXX American Chemical Society

In prior work, we have shown how these OTEs can be used


in transmission spectroelectrochemical measurements of ferri/
ferrocyanide,1,6 ferrocene,3 chlorpromazine,4 and cytochrome
c6,8 in the UVvis region and ferri/ferrocyanide in the mid- to
far-IR region.7 We have also compared the properties of BDD/
quartz OTEs with traditional indium-doped tin oxide (ITO)
electrodes.5 Several conclusions were drawn from this work.
First, diamond thin lms on quartz can be prepared with
reproducible electrical, optical, and electrochemical properties.
Second, the electrical and optical properties of diamond are
stable during exposure to a wide range of organic solvents and
aggressive aqueous solutions. In contrast, severe lm
degradation of ITO occurs during exposure to dichloromethane, 1 mol L1 HNO3, and 1 mol L1 NaOH. Third,
diamond is not as electrically conducting as ITO (ca. 102 vs
105 cm) and is not as optically transparent (55 vs 85%
transparency in the visible region). However, these are not
drawbacks for dierence spectroelectrochemical measurements
(in a thin-layer cell) as long as relatively low scan rates (<100
mV s1) and low analyte concentrations (1 mmol L1) are
employed. Fourth, the electrical, optical, and electrochemical
properties of diamond are stable during both anodic and
cathodic polarization in 1 mol L1 HNO3 and 1 mol L1 NaOH
at current densities of up to 5 mA cm2. On the other hand,
Special Issue: Electrochemical Applications of Carbon Nanomaterials
and Interfaces
Received: February 27, 2016
Accepted: May 13, 2016

DOI: 10.1021/acsami.6b02467
ACS Appl. Mater. Interfaces XXXX, XXX, XXXXXX

Forum Article

ACS Applied Materials & Interfaces

end of the deposition period, the CH4 and B2H6/H2 gas ows were
stopped, and the diamond lm remained exposed to the H2 plasma
(196 sccm). The lm was then cooled in the H2 plasma over a period
of 2030 min by slowly reducing the power and pressure down to 150
W and 10 Torr. This post-treatment cooled the sample to an estimated
temperature of less than 400 C in the presence of atomic
hydrogen;19,20 this step is critical for minimizing sp2 carbon impurity
incorporation and for maintaining stable H surface termination.
2.2. Atomic Force Microscopy. AFM was performed with a
NanoScope IIIa scanning probe microscope (Veeco Instruments Inc.,
CA) operating in the tapping mode. Pyramidal-shaped Si3N4 tips
mounted on gold cantilevers (100 m legs, 0.38 N m1 spring
constant) were used in air to obtain the topographical images.
2.3. UVVis Transmission Measurements. Optical spectroscopy was performed after the lm was rinsed with ultrapure water and
ultrapure isopropanol (distilled and stored over activated carbon) and
dried under a stream of N2. The BDD/quartz sample was placed in the
optical beam of a commercial UVvis spectrophotometer (Shimadzu
Model UV2401-PC). A piece of clean quartz was placed in the
reference beam of the instrument. Transmission spectra were recorded
in the range of 200900 nm. The refractive indices of diamond,
quartz, and air are 2.41, 1.54, and 1.00, respectively.
2.4. Raman Spectroscopy. Raman spectroscopy was performed
using a Renishaw inVia Reex spectrometer. The instrument consisted
of a confocal microscope connected to a continuous-wave diodepumped solid-state laser with a fundamental emission at 532 nm. The
sample was positioned under the laser light using a motorized stage.
The stage position was controlled and the spectral data were acquired
with the commercial WiREInterface software. This software allows for
control over the laser power, exposure time, spectral range covered,
and stage positioning for mapping of a sample. The parameters used
were a laser power of 10 mW at the sample and an integration time of
10 s. Each spectrum was generated from an average of ve spectral
acquisitions at each point. A Leica (100/0.85) objective lens was
used to focus the excitation light and collect the scattered radiation. A
1800 lines mm1 holographic grating was used, and the step size in the
mapping was 25 m.
2.5. X-ray Photoelectron Spectroscopy. XPS was performed
courtesy of Kratos Analytical Ltd. using an AXIS Supra instrument.
The surface science station was equipped with XPS and an Ar+ gas
cluster ion source for depth proling. The ion source was operated at 4
kV for depth proling of the diamond lms. X-rays were generated
from an Al/Mg K X-ray source, and the emitted photoelectrons were
analyzed using a hemispherical analyzer. The X-ray power was 600 W,
and scans were acquired over a 700 m 300 m area. The pass
energy was 160 eV, and the acquisition time was 120 s per survey scan.
The samples were analyzed as received after evacuation to the base
pressure of ca. 1010 Torr. The actual line shape used to t the spectra
was a combined Gaussian/Lorentzian shape. There was a 30%
Lorentzian component in the peak shapethis approximates the
natural line shape of the emitted electrons from within the sample.
The low-binding-energy side of the peak was t rst by adjusting the
height and half-width. Once that was done, additional components
were added on the basis of knowledge of the dierent types of
carbonoxygen functional groups and their respective binding energy
shifts from the C 1s base peak.
2.6. Hall Eect Measurements. A commercially available cryostat
was modied by the addition of a custom-designed sample stage
consisting of an aluminum nitride plate on a heated copper stage.21
The measurement temperatures were from room temperature to 700
K with the samples held under vacuum. Electrical connection to the
sample was made using 25 m diameter gold wires connected to Ti
metal contacts formed on the sample surface by physical vapor
deposition (PVD). For all of the measurements reported herein, a
four-point van der Pauw contact arrangement was employed.
The BDD sample was cleaned prior to PVD of the Ti contacts near
the four corners of the sample. The sample was then oxygen-plasmatreated to remove any surface conductivity eects. The sample was
mounted in the Hall eect system and annealed at 700 K under
vacuum while the IV behavior of the metal contacts was monitored

severe degradation of ITO occurs during polarization in either


of these media. The degradation mostly occurs during cathodic
polarization. Conducting-probe atomic force microscopy (CPAFM) measurements conrmed the microstructural stability of
diamond and the localized degradation of ITO.
There have been reports of diamond lm growth on quartz
substrates, either plates or optical bers. For example, the
growth of nanocrystalline911 and ultrananocrystalline12
diamond thin lms on quartz has been reported. A focus of
these studies was understanding how the preparation of the
quartz substrate by ultrasonic seeding with dierent-sized
diamond powders and the CVD deposition conditions aected
the lm morphology and optical properties. A key parameter
for achieving a thin and continuous lm across the surface is the
initial or primary nucleation density on the substrate. Hikavyy
et al.13 and Gajewski et al.14 reported on the structural and
electrical properties of nanocrystalline BDD lms deposited on
quartz as a function of the boron doping level. Zhang et al.15
reported on the fabrication and characterization of a thin BDD
lm deposited on a quartz oscillator for a quartz crystal
microbalance. Kulesza16 described the growth of BDD lms
deposited at lower temperatures (ca. 500 C). The objective of
their work was to maximize the optical transparency in the
visible and near-IR regions of the spectrum. Henychova et al.17
studied the eect of the deposition conditions on the properties
of nanocrystalline BDD lms on quartz as possible implantable
electrodes. Finally, Bogdanowicz et al.18 reported on the
formation of nanocrystalline BDD lms on a fused silica optical
ber.
In this paper, we report a comprehensive investigation of the
morphological (AFM), optical (transmission spectroscopy),
microstructural (Raman imaging), electrical (Hall Eect
measurements), chemical (X-ray photoelectron spectroscopy
(XPS) depth proling), and electrochemical properties of a
BDD/quartz OTE. Electrochemical data are comparatively
presented for aqueous and ionic liquid media. This work is a
continuation of our eorts to fully understand the relationship
between surface structure and the electron-transfer reactivity of
diamond OTEs.

2. METHODS AND MATERIALS


2.1. Nanocrystalline BDD Film Growth. A 1 cm2 piece of quartz
glass was rst ultrasonically cleaned in acetone for 20 min. The quartz
was very smooth, with an AFM-determined roughness of 1.2 0.2
nm. The cleaned quartz was then mechanically polished with 100 nm
diameter diamond powder (Tomei Diamond) suspended in ultrapure
water. The polishing was performed by hand for 3 min on a dedicated
felt polishing pad. After the polishing, the quartz substrate was rinsed
with a stream of ultrapure water and then ultrasonically cleaned for 30
min in acetone to remove polishing debris. This cleaning is a critical
step in the substrate surface preparation process.
The mechanical polishing step was followed by ultrasonic seeding in
Opal Seed (Adamas Nanotechnologies Inc., Raleigh, NC) for 30
min. This suspension contains 45 nm diameter primary detonation
nanodiamond particles and nominally 30 nm aggregates, all in
dimethyl sulfoxide (DMSO), as specied by the supplier. We
performed no particle size analysis on the dispersion prior to use.
Afterward, the substrate was rinsed with ultrapure water and then
wicked dry with a Kimwipe and placed in the CVD reactor for an
overnight pumpdown.
The growth of the BDD lm was performed by microwave-assisted
CVD using a commercial 1.5 kW reactor (Seki Technotron). The
growth was performed for 23 h using the following conditions: 600
W power, a total gas ow of 200 sccm (2.00 sccm CH4 + 2.00 sccm
B2H6/H2 + 196 sccm H2), and a system pressure of 25 Torr. At the
B

DOI: 10.1021/acsami.6b02467
ACS Appl. Mater. Interfaces XXXX, XXX, XXXXXX

Forum Article

ACS Applied Materials & Interfaces

Figure 1. Tapping-mode AFM images (500 nm 500 nm) of (A) a piece of quartz that was mechanically polished (100 nm diameter diamond grit
for 3 min) and ultrasonically seeded (45 nm detonation nanodiamond with 30 nm aggregates in DMSO for 30 min) and (B) a nanocrystalline
diamond overlayer formed after 30 min of growth on the mechanically polished and ultrasonically seeded quartz.
from room temperature to 700 K. Once the linear IV behavior was
observed for all contact pairs, resistivity and Hall eect measurements
were recorded using the ASTM standard procedure.22
2.7. Electrochemical Characterization. Electrochemical measurements were made using a computer-controlled workstation (model
650A, CH Instruments, Inc., Austin, TX) in a standard three-electrode
glass electrochemical cell.23 The BDD/quartz electrode was clamped
to the bottom of the cell. A Viton O-ring (i.d. 0.5 cm) placed between
the cell opening and the electrode surface ensured that a reproducible
area of 0.2 cm2 was exposed to the electrolyte solution. Electrical
contact with the diamond surface was made using a piece of aluminum
foil pressed along the entire width of the sample and a metal clip. The
O-ring was ultrasonically cleaned in ultrapure water and rinsed with
ultrapure isopropanol prior to use. Once mounted in the cell, the BDD
electrode was cleaned by soaking in ultrapure isopropanol (distilled
and stored over activated carbon) for 20 min. The cell was then rinsed
with the supporting electrolyte. Afterward, the cell was lled with the
analyte solution of interest. The counter electrode was a coiled Pt wire
that was positioned normal to the working electrode. The reference
electrode was a commercial Ag/AgCl electrode lled with 3 mol L1
KCl saturated in AgCl (0.197 V vs NHE, Bioanalytical Systems, West
Lafayette, IN).
2.8. Capacitance Measurements. Dierential capacitance
measurements were made using a potentiostat (PAR model 173)
and a lock-in amplier (Stanford Research Systems). A 10 mV
alternating-current (ac) sine wave was coadded to the applied directcurrent (dc) potential, and the capacitance was determined from the
following equation:

Z im =

pseudoreference electrode (0.128 V vs Ag/AgCl measured in


[EMIM][BF4]).
2.9. Ionic Liquids and Purication. The room-temperature ionic
liquid (RTIL) used in this work, 1-butyl-3-methylimidazolium
tetrauoroborate ([BMIM][BF4]), was purchased from Sigma-Aldrich
(97.0%; cat. no. 91508). The dielectric constant of the RTIL was
14.5.24 The RTIL viscosity, , is typically 10100 times greater than
that of water. [BMIM][BF4] has a reported viscosity of 112 cP.24
The following procedure was used to remove trace organic and
water contamination from the RTIL.25,26 The as-received RTIL was
rst stored over activated carbon for 3 days. After this period, the
liquid was centrifuged to settle out the carbon powder. Most of the
RTIL sample (several milliliters) was then carefully removed and
stored. At the time of an electrochemical measurement, a small volume
(12 mL) of this RTIL was transferred to the electrochemical cell.
The transfer was performed in a N2-lled glovebox (Coy Laboratories,
Grass Lake, MI). Once in the electrochemical cell, the RTIL was
heated at 70 C for 50 min while being purged with ultrapure Ar
(99.995%, Linde). An advantage of this sweeping purication
method is the lower water level that can be achieved compared with
the vacuum drying method. Another advantage of sweeping
purication is the short treatment time (50 min vs multiple hours or
days for vacuum drying). Additionally, the sweeping method is easily
performed directly in the electrochemical cell prior to a measurement,
which eliminates the possibility of water contamination during transfer
of the RTIL from the vacuum apparatus to the glovebox.

3. RESULTS
3.1. Substrate Seeding. Preparing the quartz surface
properly ahead of the diamond growth is a critical step for
achieving a thin and uniform lm. A uniform and high initial
nucleation density is needed to grow a thin and continuous
nanocrystalline diamond lm in a short time. This is
accomplished by mechanical polishing and ultrasonic seeding
of the quartz with nanoparticles of diamond. The mechanical
polishing step introduces striations (trenches) on the substrate
surface, as can be seen in the AFM image presented in Figure
1A. Some residual carbon material likely remains in the
striations even after the ultrasonic cleaning that follows the
polishing. Analysis of the AFM images revealed a nominal
striation width of 42 9 nm. The apparent depth of the
striations was only about 2 nm. The clean scratches and
adventitiously embedded diamond/carbon particles serve as the
initial nucleation sites for diamond growth.

1
2fCdl

where Zim is the imaginary component of the total impedance, f is the


ac frequency used (in Hz), and Cdl is the capacitance at the particular
potential (E). The measurements were made using a single frequency
of 10 Hz. This frequency was chosen on the basis of experimental
results that revealed the measured capacitance of ca. 5 F cm2 for a
BDD thin lm in 1 mol L1 KCl at 0.1 V vs Ag/AgCl (3 mol L1 KCl)
to be independent of the frequency between 0.1 and 103 Hz.
Additionally, the low frequency was used to allow sucient time for
the ions, especially in the ionic liquids, to reorganize in response to the
potential changes. Ionic liquids are considerably more viscous that
water, and therefore, they exhibit more sluggish transport kinetics.
Measurements in the aqueous electrolyte made use of the previously
mentioned commercial Ag/AgCl reference electrode, while the
measurements in the two ionic liquids made use of a Ag wire
C

DOI: 10.1021/acsami.6b02467
ACS Appl. Mater. Interfaces XXXX, XXX, XXXXXX

Forum Article

ACS Applied Materials & Interfaces


The tapping-mode AFM image in Figure 1A reveals diamond
particles distributed over the quartz substrate after the seeding.
Some of these particles reside in the polishing striations, while
most are positioned on the quartz surface outside of these
scratches, as indicated by the bright yellow features in the
image. The particle sizes range from less than 10 nm to as large
as 75 nm. Many were found to have diameters in the 2040 nm
range, consistent with the 30 nm nominal aggregate sizes in the
suspension. Counts of the visible particles on the quartz surface
after the seeding pretreatment revealed a seed density of 1010
cm2. This surface preparation indeed produces a high and
uniform initial nucleation density, as a continuous nanocrystalline BDD lm is seen over the surface (500 nm 500 nm) in
Figure 1B after a short (30 min) diamond deposition. The
diamond crystallites have a narrow size dimension in the 1040
nm range.
3.2. Film Morphology. Figure 2 shows tapping-mode AFM
images of a nanocrystalline BDD lm deposited on quartz after

Figure 3. (A) Transmission spectrum and (B) optical image of a BDD


OTE. The nanocrystalline BDD thin lm was ca. 0.51 m thick and
continuous over the quartz substrate surface.

the same type of nanocrystalline BDD lm deposited on


quartz.4 The transmission spectrum shows a low-wavelength
cuto at ca. 220 nm due to the band-gap absorption by
diamond. A relatively constant transmittance of 5060% is seen
between 400 and 700 nm. The transmittance decreases at
longer wavelengths above 700 nm. The majority of the light
loss for diamond, particularly at wavelengths from 220 to 300
nm and greater than 700 nm, is attributed to absorption by an
adventitious nitrogen impurity in the lm and the boron
impurity band, respectively. The nitrogen arises from air
leakage into the reactor during lm growth, as the systems base
pressure is only ca. 10 mTorr. The band-gap absorption edge is
at 220 nm (5.4 eV). The lm is bluish as a result of absorption
of red light by the boron impurity band. Reection is a main
light-loss mechanism in the visible region because of diamonds
relatively high refractive index (2.41 at 590 nm).4,5 The
refractive index of quartz is 1.54, and of course, that of air is
1.00. The calculated intensity loss due to reection (Snells law)
for light incident on the BDD/quartz OTE is ca. 27%.
Importantly, prior research has shown that the optical and
electrical properties are stable over a wide potential range,
especially at negative potentials, and during exposure to a
variety of chemical environments.5 This stability is essential for
OTEs used in dierence transmission spectroelectrochemical
measurements.1,37,27
3.4. Film Microstructure. Raman imaging spectroscopy
was used to probe the lm microstructure spatially. Shown in
Figure 4 are a series of 40 spectra recorded over an area in the
center of the lm, along with a representative spectrum from
the series. Similar spectra were acquired in the outer four
quadrants of the sample. Multiple peaks are seen in the
spectrum at 488, 1137, 1215, 1320, 1470, and 1550 cm1.
These spectral features are retained over the entire mapped
area. This indicates that the diamond microstructure is uniform
across the lm. Importantly, these spectral features are
consistent with the lm being heavily boron-doped.28,29 For
heavily boron-doped lms, the rst-order diamond phonon line
shifts from 1332 cm1 to lower wavenumbers with increasing
boron concentration. This shift is typically accompanied by
increases in scattering intensity at ca. 500 and ca. 1225
cm1.28,29 In these spectra, the peak is shifted from the expected
1332 cm1 position down to 1320 cm1. There is also
signicant scattering intensity at 488 and 1215 cm1. The
scattering at 488 cm1 dominates the spectrum at all positions.
Because the intensity of the 488 cm1 band increases with
boron doping level, this peak has been assigned to the
vibrational modes of boron dimers3032 and pairs or clusters.33
The 1215 cm1 peak has been assigned to defects in the

Figure 2. Tapping-mode AFM images (500 nm 500 nm) of a


nanocrystalline BDD thin lm deposited on quartz: (A) line scan; (B)
plan view. The growth time was 2 h.

the full growth time of 2 h. The lm thickness was 0.5 to 1 m,


as estimated from side-view scanning electron microscopy
images of the lm (data not shown) and from mass gain
measurements after the growth. With a density of 3.51 g cm3
for diamond, the apparent lm thickness l (in cm) was roughly
estimated from the following equation:
l=

w
dA

in which w is the mass gain (in g), d is the density (in g


cm3), and A is the area of the lm (in cm2). The height-mode
AFM image in Figure 2A reveals a nanocrystalline morphology,
with primary grains whose diameters are between 50 and 100
nm. There is also evidence of secondary nucleation, as smaller
growths can be seen on the primary grains and in the grain
boundaries. This grain size, within order of magnitude, is
similar to the grain size after the 30 min growth (Figure 1B).
There is little grain coarsening during the short deposition
period. The plan view image in Figure 2B reveals a mixture of
primary grains, some of which are labeled (6085 nm
diameter), and smaller secondary growths. These images are
consistent with prior AFM data reported for a diamond/quartz
OTE.4,5
3.3. Optical Properties. For spectroelectrochemical
measurements, the optical transparency and electrical conductivity and the stability of these two properties with potential
are critical. Figure 3A presents a typical transmission spectrum
for a nanocrystalline BDD thin lm deposited on quartz. Figure
3B shows an optical image of a BDD OTE placed over a ruler.
These data are consistent with previously reported results for
D

DOI: 10.1021/acsami.6b02467
ACS Appl. Mater. Interfaces XXXX, XXX, XXXXXX

Forum Article

ACS Applied Materials & Interfaces

Figure 4. (left) Series of Raman spectra recorded along a line prole across a nanocrystalline BDD thin lm on quartz. (right) Individual spectrum
showing peaks at 488, 1137, 1215, 1320, and 1470 cm1 and a shoulder at 1550 cm1. The 40 spectra were recorded over an area of 116 m 179
m. Therefore, each spectrum was recorded about every 4.5 m along the long axis.

diamond lattice brought about by the high boron doping,


possibly boroncarbon complexes.28,29 The 1137 cm1 peak
and its companion peak at 1470 cm1 have been assigned to
polymeric sp2 carbon species in the grain boundaries.34 Finally,
the 1550 cm1 shoulder is caused by some crystalline graphite
that becomes more prominent with increasing boron
concentration.28,29 In other words, the addition of large
amounts of boron is accompanied by increasing carbon
graphitization in the grain boundaries. Similar spectral maps
were observed in dierent quadrants of this particular lm and
on other BDD OTEs. Even with the crystalline graphite
impurity phase, the electrochemistry of the BDD lm is
controlled by the diamond and not the sp2 carbon, which is
low in abundance.
3.5. Electronic Properties. Hall eect measurements were
used to evaluate the electronic properties of the BDD OTE in
terms of electrical resistivity, carrier concentration, and carrier
mobility as functions of temperature. The measurements were
made in a four-contact van der Pauw arrangement using Au/Ti
contacts. The measurements were made at temperatures of
300500 and 300700 K. Figure 5 shows plots of the
resistivity as a function of temperature. The initial measurements were made at increasing temperature from 300 to 500 K
(black data). The curve reveals a slight decrease in resistivity
with increasing temperature. The room-temperature resistivity
of the diamond lm is ca. 0.06 cm. The measurements were
then repeated from 300 to 700 K. Again, there is a very small
factor of 2 decrease in the resistivity over the temperature range
during the second temperature ramp. Overall though, there is
little dierence in the resistivities for these two temperature
ramps. The similar results for the two heating cycles indicate
that good ohmic contact was achieved initially with the Au/Ti
contacts and that exposure to the higher temperatures did not
signicantly reduce the contact resistance through annealing.
The weak temperature dependence of the resistivity is
consistent with the diamond lm being a degenerate semiconductor/semimetal. It should be noted that the resistivity
decreases from 0.06 to only 0.035 cm (i.e., the conductivity
increases) over a 400 K temperature range. Generally, the
temperature dependence of the electrical conductivity has a
competition mechanism: the number of activated carriers
increases with increasing temperature, but their mobility

Figure 5. Plots of the resistivity as a function of temperature for a


nanocrystalline BDD thin lm on quartz. Two sets of data are
presented: (i) for a temperature range of 300500 K (black data) and
(ii) for a temperature range of 300700 K (red data). The narrow
range of the y axis should be noted. There was no hysteresis in the
resistivity values with increasing then decreasing temperature, as the
values at each temperature were identical.

decreases. For the BDD OTE, these two eects largely oset
each other, leading to the observed weak temperature
dependence. In the case of a degenerate semiconductor, there
are many carriers readily available for conduction so large
changes with increasing temperature are not expected. This is
further evidenced in the data presented below.
Figure 6 shows plots of (A) the charge carrier concentration
and (B) the carrier mobility as functions of temperature for the
BDD OTE. Generally, consistent with the weak temperature
dependence of the resistivity, both the carrier concentration
and the mobility are largely independent of the temperature.
Qualitatively, there is a slight increase in the carrier
concentration with temperature. This trend reects a slight
increase in the number of thermally activated carriers. Overall,
however, these electrical measurements are consistent with the
thin nanocrystalline BDD lm being a degenerate semiconductor/semimetal. The carrier concentration is high (1021
cm3) because of the high doping level, and the mobility is
relatively low (0.2 cm2 V1 s1) because of carrier scattering by
E

DOI: 10.1021/acsami.6b02467
ACS Appl. Mater. Interfaces XXXX, XXX, XXXXXX

Forum Article

ACS Applied Materials & Interfaces

Figure 6. Plots of (A) the charge carrier concentration and (B) the charge carrier mobility as functions of temperature for a nanocrystalline BDD
thin lm on quartz. Measurements were made over a temperature range of 300700 K. In (A), the carrier concentration at 410 C is likely a
measurement anomaly and is clearly inconsistent with the other data.

Figure 7. XPS survey spectrum for a nanocrystalline BDD thin lm on quartz (A) over a broad energy range from 0 to 1195 eV and (B) over a
narrow range from 0 to 300 eV. The nanocrystalline diamond lm thickness was 0.51 m. The OTE was exposed to the laboratory air for multiple
weeks prior to the measurement.

Figure 8. XPS depth proles through a nanocrystalline BDD thin lm deposited on quartz. The depth proling was performed using a 4 kV Ar+
beam. The depth proles on the left include signals for C, O, Si, and Ar. The depth proles on the right show the B and Ar signals (in the blueshaded region) on an expanded scale; signals for C (red), O (blue), and Si (green) are o-scale.

lm and to examine the chemical environment around the C


and B atoms. Figure 7 shows survey scan data for the BDD
OTE. The elements present, C and B, are expected, as is a low
level of surface oxygen due to the laboratory air exposure. Their
relative atomic concentrations are 96.48, 0.51, and 2.94%,
respectively. Signals for Si and S are also present, although at

the high fraction of grain boundaries resulting from the


nanometer-sized grains. The high carrier concentration is
consistent with the high doping level, as reected in the Raman
spectra.28,29,35,36
3.6. Chemical Properties. XPS was used to measure the
depth prole of the boron concentration through the diamond
F

DOI: 10.1021/acsami.6b02467
ACS Appl. Mater. Interfaces XXXX, XXX, XXXXXX

Forum Article

ACS Applied Materials & Interfaces

Figure 9. Peak ts of the (A) C 1s and (B) B 1s regions of the XPS survey spectrum for a nanocrystalline BDD thin lm deposited on quartz. The
results reveal the dierent chemical environments around the C and B atoms in the lm.

Figure 10. Cyclic voltammetric jE curves for 0.1 mmol L1 analytes in 1 mol L1 KCl at a nanocrystalline BDD thin lm on quartz. The redox
analytes are (A) Fe(CN)63/4, (B) Ru(NH3)63+/2+, (C) IrCl62/3, and (D) methyl viologen. Scan rate = 0.1 V s1. Arrows depict the direction of
the initial scan. The lm thickness was 0.51 m.

very low levels (0.17 and 0.04%, respectively). These two


elements arise from the quartz substrate. Even though this
particular lm was exposed to the atmosphere for multiple
weeks prior to the XPS measurements, the surface is low in
oxygen content as the atomic O/C ratio is 0.03.
Figure 8 shows a set of XPS depth prole data for several
elements through the diamond lm. The proles are presented
as the relative atomic concentration (%) versus the sputtering
time (i.e., depth). The sputtering was accomplished using a 4
kV Ar+ beam. Elemental signals for C 1s, O 1s, and B 1s are
shown. Once the lm is sputtered through and the quartz
substrate is reached, the C 1s signal decreases while the O 1s
and Si 2p signals increase in intensity. Importantly, the depth
prole shown on an expanded scale (right) reveals a relatively
constant B 1s signal with depth. This indicates that the boron
doping is uniform at ca. 0.5 atom % with the diamond lm

thickness. This corresponds to 5000 ppm B/C or a boron


concentration of ca. 9 1020 cm3. This high doping level is
consistent with the Raman spectral data and the high carrier
concentration revealed by the Hall eect measurements. The
Ar 2p signal results from the Ar ion beam used for sputtering.
Figure 9A,B shows deconvoluted XPS spectra from the C 1s
and B 1s regions, respectively. Deconvolution of the C 1s peak
in Figure 9A reveals the dominant signal at 285.0 eV, which is
assigned to sp3 C. The weak-intensity peak at 283.6 eV is
assigned to non-diamond sp2 C impurity in the lm. This
results from the high boron concentration, which causes some
sp2-bonded carbon formation. The presence of a small amount
of crystalline graphite is evident in the Raman spectral data (see
Figure 4). Clearly, though, the sp2 impurity content is very low.
Two higher-binding-energy components are also seen at 286.0
eV (+1.0 eV) and 287.2 eV (+2.2 eV). On the basis of the
G

DOI: 10.1021/acsami.6b02467
ACS Appl. Mater. Interfaces XXXX, XXX, XXXXXX

Forum Article

ACS Applied Materials & Interfaces

Figure 11. Cyclic voltammetric jE curves for 0.1 mmol L1 analyte in 1 mol L1 KCl at a BDD OTE. The redox analytes are (A) Fe(CN)63/4,
(B) Ru(NH3)63+/2+, (C) IrCl62/3, and (D) methyl viologen. Curves are shown for each analyte as a function of scan rate (0.1, 0.2, 0.3, 0.4, and 0.5
V s1). The lm thickness was 0.51 m.

binding energy shift, these are the oxide components and


result from ether (COC) or hydroxyl (COH) and
carbonyl (CO) functional groups, respectively.3740
As discussed above, the O/C atomic ratio in the lms nearsurface region is low, only ca. 0.03. The deconvoluted B 1s
spectrum shown in Figure 9B reects three distinct chemical
environments for the boron in the diamond lm. These three
components are at 186.5, 187.7, and 193.0 eV and are assigned
to BB, BC, and BO bonding states in the lm,
respectively.41,42 The atomic contents of these states increase
as BO (193.0 eV) < BB (186.5 eV) < BC (187.7 eV).
Clearly, most of the boron in the lm is substitutionally
inserted into the diamond lattice (BC).
3.7. Electrochemical Properties. The above data provide
a clear picture of the physical, chemical, and electronic
properties of the BDD OTE. Cyclic voltammetry along with
several redox probes was used to evaluate the basic electrochemical properties of the BDD OTE. The redox probes were
selected on the basis of their wide range of standard reduction
potentials. Figure 10 shows cyclic voltammograms (v = 0.1 V
s1) for (A) Fe(CN)63/4, (B) Ru(NH3)63+/2+, (C) IrCl62/3,
and (D) methyl viologen (MV). The concentration of all was
0.1 mmol L1, and the supporting electrolyte was 1 mol L1
KCl. The total current curve and the background curve are
shown for each redox system. Nearly reversible behavior is seen
for all four redox probes at this scan rate. Well-dened
oxidation and reduction peaks are seen for each. All of the
reactions were under diusion control, as evidenced by the fact
that the forward peak current for each probe molecule varied
linearly with v1/2, where v is the scan rate (see below). The peak
currents also increased linearly with the analyte concentration
from 0.1 to 1 mmol L1 (data not shown). The background
currents in 1 mol L1 KCl were at and devoid of any peaks in
the potential region where the redox probes undergo electron
transfer. For methyl viologen, there are two redox peaks seen at
0.65 V vs Ag/AgCl for the MV2+/MV+ transition and at 1.0
V vs Ag/AgCl for the MV+/MV transition. These results

indicate that the BDD OTE exhibits a high level of redox


activity without conventional pretreatment (e.g., mechanical
polishing + electrochemical pretreatment). In fact, the only
pretreatment used was a 20 min soak in ultrapure isopropanol.
The heterogeneous electron-transfer rate constants for Ru(NH3)63+/2+, IrCl62/3, and methyl viologen on carbon
electrodes, including BDD, are most sensitive to the electronic
properties of the electrode (i.e., density of electronic states)
around the E of the redox probe.40,4352 The relatively large
rate constants indicate that the BDD OTE possesses a high
density of electronic states over a wide potential range from
1.2 to 1.0 V vs Ag/AgCl. This is reective of the semimetallic
nature of the diamond, consistent with the Hall eect
measurement data presented in Figure 6. The rate constant
for Fe(CN)63/4 is quite sensitive to the surface cleanliness
and surface chemistry of carbon electrodes as well as the
electronic properties.40,4352 On BDD, the rate constant
decreases with increasing surface carbonoxygen content.40,5052 The relatively small Ep for this redox system
(ca. 100 mV) indicates that the BDD OTE has a clean surface
devoid of signicant levels of surface oxygen (ca. 3 atom % in
the XPS data).
Figure 11AD shows cyclic voltammograms for (A)
Fe(CN)63/4, (B) Ru(NH3)63+/2+, (C) IrCl62/3 and (D)
methyl viologen as a function of the scan rate. The
concentration of all was 0.1 mmol L1 and the supporting
electrolyte was 1 mol L1 KCl. In all cases, the peak currents
increase with the scan rate. The forward peak current for each
redox analyte increased linearly with v1/2 (r2 > 0.98), conrming
that the redox reactions are all diusion-controlled. Ep
increased with increasing scan rate for all of the redox systems,
consistent with quasi-reversible electron-transfer kinetics. Ep
ranged from 125 to 158 mV for Fe(CN)63/4, 96 to 105 mV
for Ru(NH3)63+/2+, 62 to 72 mV for IrCl62/3, and 6070 mV
for methyl viologen over the scan rate range from 50 to 500 mV
s1.
H

DOI: 10.1021/acsami.6b02467
ACS Appl. Mater. Interfaces XXXX, XXX, XXXXXX

Forum Article

ACS Applied Materials & Interfaces

Figure 12. (A) Background cyclic voltammetric scans as a function of scan rate for a BDD OTE in 0.5 mol L1 H2SO4. Scan rates from 0.1 to 0.5 V
s1 were used. (B) Plot of log current at 0.3 V vs log scan rate.

Figure 13. Capacitancepotential curves for a nanocrystalline BDD thin lm on quartz. The lm thickness was 0.51 m. Curves are shown for (A)
0.5 mol L1 H2SO4 and (B) [BMIM][BF4]. Capacitance values were determined at a single frequency of 10 Hz. Cdl was calculated from the following
1
equation: Z im = 2fC .
dl

3.8. CapacitancePotential Proles. Finally, the capacitance of the BDD OTE was investigated as a function of
potential in two dierent electrolyte media: aqueous electrolyte
and ionic liquid. Figure 12A shows background cyclic
voltammetric scans as a function of scan rate for the BDD
OTE in 0.5 mol L1 H2SO4. The background current increases
as the scan rate is increased from 0.1 to 0.5 V s1. Figure 12B
shows a loglog plot of the background current (A cm2) at
0.3 V versus the scan rate (V s1). A linear relationship is seen,
with a slope of 0.89. A slope of 1.0 is expected for this plot if
the background current is capacitive in nature. Thus, the
background current for the diamond OTE in this potential
region is capacitive. The capacitance can be calculated from the
slope of the plot, and thus Cdl at 0.3 V was found to be 7.7 F
cm2.
Figure 13 shows CdlE curves for the electrode in (A) 0.5
mol L1 H2SO4 and (B) [BMIM][BF4]. The capacitance values
were measured at a single frequency (10 Hz). The capacitance
in the H2SO4 aqueous solution ranges from 47 F cm2 over
the potential range probed, with little hysteresis observed
during the potential cycle. The capacitance exhibits a general
trend of increasing magnitude with increasing positive
potential. The capacitance measured by the ac method is
consistent with the value determined by cyclic voltammetry
within an order of magnitude. The low capacitance value is

typical of BDD and has been reported since the very early days
of studies of the material.53,54 The relatively low capacitance, as
compared with glassy carbon for example (2530 F cm2), is
due to two factors: (i) a lower potential-dependent density of
states due to the semimetal electronic properties (i.e., less
excess surface charge at the dierent potentials) and (ii) the
relative absence of electroactive and ionizable surface carbon
oxygen functional groups that lead to pseudocapacitance on
graphitic carbon electrodes.54,55 The data reveal that there is
little hysteresis in the capacitance recorded during positive- and
negative-going applied potentials. This reects the stability of
the diamond electrode surface chemistry and microstructure at
these potentials.
The capacitance in the ionic liquid ranges from 0.7 to 1 F
cm2. The lower capacitance, to a rst approximation, is
consistent with the lower dielectric constant of the ionic liquid
(14.524) compared with that of water. The capacitance
potential curves in the RTIL also exhibited little hysteresis,
which contrasts with observations previously reported.56 This
indicates that the surface structure (chemistry and microstructure) of the diamond lm and that of the ionic liquid are
stable at these potentials.
I

DOI: 10.1021/acsami.6b02467
ACS Appl. Mater. Interfaces XXXX, XXX, XXXXXX

Forum Article

ACS Applied Materials & Interfaces

4. DISCUSSION
The results reported herein provide the rst comprehensive
investigation of the relationship between the physical and
chemical structure and electronic properties of a diamond OTE
and the electrodes electrochemical activity. With any carbon
electrode, it is critical to have a good understanding of the
surface chemistry and microstructure as well as the materials
electronic properties if one is going to reproducibly achieve
high-quality electroanalytical measurements. This means low
background current and noise, weak adsorption or fouling, and
high rates of electron transfer with redox analytes.
Key for preparing BDD OTEs is achieving a balance with
respect to both the nominal lm thickness required for
achieving a continuous lm and the doping level used for
imparting electrical conductivity. Ideally, one desires a lm that
is as thin as possible but continuous and the minimum doping
level needed to impart sucient electrical conductivity for the
desired measurements. Non-diamond substrates require some
pretreatment in order to enhance the nucleation density to the
point where very thin coalesced lms can be grown. This is
certainly the case for quartz. In our hands, this is reproducibly
accomplished by a combination of careful mechanical polishing
and surface cleaning with ultrasonic seeding in a nanodiamond
powder suspension. Both the polishing striations, likely the
carbon residue in the striations, and the nanodiamond particles
that decorate the surface serve as the initial nucleation sites.
Our substrate preparation protocol produces a high seed
particle density (ca. 1010 cm2) uniformly over the surface. The
resulting nanograined diamond lm is relatively thin (0.51
m) and continuous over the surface and strongly adheres to
the quartz substrate. Others have looked carefully at the proper
ways of seeding surfaces to achieve nanocrystalline diamond
lms.9,11,12,5759
The BDD OTE exhibits ca. 60% transparency in the visible
region of the electromagnetic specrum.4,5 The main light loss
mechanism is due to reection because of the high refractive
index of diamond (2.41). The reduced light throughput at
wavelengths below 250 nm is due to the band-gap absorption
of diamond (5.5 eV).4,5 The reduced transparency in the nearIR region is due to absorption by boron impurity states in the
doped lm.4,5 The absorption in the red gives the lm its blue
color. Importantly for dierence spectroelectrochemical measurements, the electrical, optical, and electrochemical properties
of the BDD OTE are stable at extreme positive and negative
potentials, during exposure to strong acid and base, and during
exposure to dierent organic solvents.5
The Raman spectroscopy data are reective of the nanocrystalline morphology of the lm, the relatively high fraction of
grain boundaries, and the substitutional insertion and possible
clustering of boron. Additionally, the spectral features are
characteristic of a heavily doped lm.35,36 The Raman mapping
reveals that the microstructure of the BDD lm is uniform
across the quartz substrate, as identical maps were obtained in
ve quadrants of the lm. The spectra presented in Figure 4
were acquired in the center of the lm, where the electrochemical measurements were performed. The grain size of the
nanocrystalline diamond determines the surface-to-volume
ratio, which is the main determining factor in the sp2 carbon
and hydrogen content.58 Smaller grains result in higher surface
areas and higher sp2 carbon and hydrogen content because of
the enhanced grain boundary volume. The diamond phonon
line is downshifted from 1332 to 1320 cm1 as a result of the

increased BC bonding within the diamond lattice due to the


substitutionally inserted boron. The intensities of the ca. 500
and 1200 cm1 bands increase with boron doping level,28,29 and
these peaks have been assigned to the vibrational modes of
boron dimers, clusters, and boroncarbon complexes.3032 The
1137 cm1 peak and its companion peak at 1470 cm1 have
been assigned to polymeric sp2 carbon species in the grain
boundaries.34 Finally, the 1550 cm1 shoulder is caused by the
in-plane stretching mode of sp2-bonded graphite-like carbon,
presumably in the grain boundaries. The sp2 carbon content
becomes more prominent with increasing boron concentration.28,29 The level of the sp2-bonded carbon is relatively low
(1.57%) on the basis of peak tting of the XPS C 1s spectral
region. Additionally, the intensities of the 1320 and 1550 cm1
modes in the Raman spectra cannot be directly related to
abundances of sp3 diamond-like and sp2 graphite-like carbon
because of the higher Raman scattering cross section for sp2
versus sp3 carbon with visible excitation.59,60 The boron-doped
nanocrystalline diamond does contain some sp2 carbon
impurity, but this impurity does not dominate the optical and
electrochemical properties.
The BDD OTE is characterized by a uniform boron
concentration with depth and a low surface carbonoxygen
functional group content (O/C = 0.03) even after lengthy
exposure to the laboratory atmosphere. The boron doping level
was constant with depth at an atomic concentration of ca. 0.5%.
This corresponds to a concentration of 5000 ppm B/C or ca. 9
1020 cm3. Deconvolution of the C 1s region of the spectrum
revealed that most of the carbon atoms in the lm are sp3hybridized (285.0 eV). Peak tting revealed that the lm
contains a low percentage (1.57%) of sp2-hybridized carbon
atoms (283.6 eV). The surface is low in oxygen content, with
two functional groups identied: single-bonded CO functional groups (286.0 eV, 8.71%) possibly as COH or CO
C, and double-bonded functional groups (287.2 eV, 1.25%),
possibly as CO carbonyl groups. Peak tting of the B 1s
region of the spectrum revealed that the boron exists in three
chemical environments: BB pairs or clusters, likely in grain
boundaries (186.5 eV, 32.5%); BC groups due to the
substitutionally inserted boron dopant atoms (187.7 eV,
55.2%); and BO groups at the lm surface (193.0 eV,
12.4%). The majority of the boron exists as BB pairs or
clusters and substitutionally inserted BC groups.
The B atomic concentration determined by XPS, the
characteristic features of the Raman spectra, and the Hall
eect measurement data all reect the high boron doping level
in the BDD OTE. An interesting nding is that the carrier
concentration determined by Hall eect measurements, 1
1021 cm3, is similar to the XPS-determined boron concentration in the lm, ca. 9 1020 cm3. This indicates that most of
the incorporated boron is electrically active. There is the
possibility that some of the carriers are associated with
hydrogen in the nanocrystalline diamond lm. For activated
conduction, the following relationship between the electrical
conductivity and the temperature is expected:61
= 0 exp(E /kT )

where E is the activation barrier for conduction. The weak


temperature dependence of the electrical resistivity (or
conductivity) and hence the carrier concentration and carrier
mobility are consistent with the diamond behaving electronically as a degenerate semiconductor or semimetal. A plot of log
versus 1000/T revealed an activation energy for conduction
J

DOI: 10.1021/acsami.6b02467
ACS Appl. Mater. Interfaces XXXX, XXX, XXXXXX

Forum Article

ACS Applied Materials & Interfaces

dipole interactions, hydrogen bonding, and -stacking forces.56


It appears that the GCS model does not adequately explain the
interfacial organization in RTILs. One model proposed is that
of Kornyshev.62 Unlike the GCS model, Kornyshevs model
predicts a bell-shaped capacitancepotential prole with a
maximum at the point of zero charge. This predicted camelshaped capacitancepotential prole has wings that are
asymmetric depending on the sizes of the anion and cation
of the RTIL. In our work, a camel-shaped prole was not
observed for the BDD OTE, as has been observed for
polycrystalline diamond lms.63 The reason that the camelshaped prole was not observed is unclear at this point. The
lower capacitance at potentials below 0 V vs Ag QRE could
arise because these potentials are negative of the point of zero
charge and the interfacial capacitance is dominated by the larger
organic cation of the RTIL. We are in the very early stages of
studying capacitancepotential relationships of BDD OTEs in
dierent RTILs and at dierent temperatures. Future work
should shed more light on the interfacial structure in RTILs.

of 10 meV. Similar low activation energies have been measured


for heavily boron-doped nanocrystalline diamond lms.58
Research has shown that the electrical conductivity increases
in nanocrystalline diamond with increasing doping level.58,61
When the doping level increases, the activation barrier for
conduction, E decreases.58,61 At high doping levels, the
conductivity becomes independent of the temperature,
reective of metallic conduction. Films with doping levels
below 1019 cm3 exhibit clear valence-band-activated transport,
with hopping transport between 1019 and 2 1020 cm3 and
nally metallic transport at higher doping levels.58 The carrier
concentration in the BDD nanocrystalline lm is sucient to
impart relatively high room-temperature conductivity (33 S
cm1) without compromising the optical throughput. In other
words, this doping level produces the appropriate compromise
between electrical conductivity and optical transparency.
The electrochemical results indicate that the low-surfaceoxygen BDD OTE exhibits a high level of activity for soluble
redox probes spanning a wide range of standard reduction
potentials, nearly 2 V.23,40,48,49,52 Quasi-reversible redox
behavior is seen for surface-sensitive (Fe(CN)63/4) and
surface-insensitive (Ru(NH3)63+/2+, IrCl62/3, and MV) redox
probes at scan rates from 100 to 500 mV s1. Importantly, the
doping level of the lm is sucient to provide a number of
charge carriers adequate to support electron transfer but not so
high as to render the lm opaque. It should be noted that the
current is measured from the center of the OTE (where
electrochemical reactions occur through the lm) to the lm
edge (where current collection is performed). The electrical
resistance through the lm is therefore inuenced by the
connectedness of the nanocrystalline grains within the diamond
lm. A highly doped lm could exhibit ohmic resistance eects
that are signicant enough to contribute to the cyclic
voltammetric Ep values if the lm is not deposited long
enough to achieve good grain coalescence. We did not perform
measurements on this particular BDD OTE to rule out ohmic
eects.
Finally, the diamond OTE is characterized by a low Cdl in
aqueous electrolyte (28 F cm2) across an extended
potential range. The low capacitance is characteristic of
diamond and is due to a lower density of electronic states in
diamond (i.e., a reduced excess surface charge density as a
function of the applied potential) compared with metal-like
glassy carbon. Also, unlike glassy carbon, there are few
electroactive and or ionizable surface carbonoxygen functionalities on BDD. This is another reason for the low capacitance.
The reduced capacitance of the BDD OTE in the RTIL is
due, to a rst approximation, to the 5.5 times lower dielectric
constant of [BMIM][BF4] (14.524) versus H2O (80). In
general, the data in Figure 13 reveal that the capacitance is 56
times lower for the RTIL at all potentials, consistent with the
dierence in dielectric constants. The electric double layer in
aqueous electrolytes is described by the GouyChapman
Stern (GCS) model.56 This model predicts that as the electrode
becomes more highly charged, the diuse layer in aqueous
electrolyte solutions becomes more compact and its dierential
capacitance increases. It is unclear whether this model
adequately describes the interfacial structure in ionic liquids.
The ions that constitute RTILs are often large, exible, highly
polarizable, and chemically complex.56 As a consequence, there
are a number of interionic forces, in addition to electrostatics,
that may act to aect the interfacial structure at an electried
interface. These forces can include dispersion forces, dipole

5. CONCLUSIONS
The following conclusions can be reached regarding the
research presented and discussed:
1. Thin and continuous diamond thin-lm optically transparent electrodes can be grown on quartz with good
adhesion and uniform lm thickness with appropriate
substrate pretreatment. The mechanical polishing and
ultrasonic seeding pretreatment described herein produces a high initial nucleation density that leads to a
continuous nanocrystalline lm without grain coarsening
in a short growth period.
2. The nanocrystalline diamond lm consists of primary
grains that are 6085 nm in diameter after 2 h of growth.
The lm exhibits ca. 60% transparency in the visible
region.
3. The microstructure of the BDD OTE is uniform across
the surface, as evidenced by the Raman mapping. The
spectral features reect the nanocrystalline morphology
of the lm, the relatively high fraction of grain
boundaries, and the substitutional insertion and clustering of boron. Additionally, the spectral features are
characteristic of a heavily boron-doped lm.
4. Temperature-dependent Hall eect measurements revealed a room-temperature resistivity of 0.06 cm that
decreases by only a factor of 2 during a temperature
ramp from 300 to 700 K. The carrier concentration (ca.
1021 cm3) and carrier mobility (ca. 0.2 cm2 V1 s1)
were largely temperature-independent, consistent with
the electronic properties of a degenerate semiconductor
or semimetal. A low activation energy of conduction of
10 meV was determined.
5. XPS revealed a surface low in carbonoxygen functionalities (O/C = 0.03). The boron doping level was
uniform with depth in the lm. The boron concentration
in the lm, ca. 1021 cm3, is consistent with the high
boron doping level reected in the Raman spectra and
with the carrier concentration determined from the Hall
eect measurements. The boron exists in three chemical
environments: BB pairs or clusters, likely in grain
boundaries; BC groups due to the substitutionally
inserted boron dopant atoms; and BO groups at the
lm surface.
K

DOI: 10.1021/acsami.6b02467
ACS Appl. Mater. Interfaces XXXX, XXX, XXXXXX

Forum Article

ACS Applied Materials & Interfaces

Spectroelectrochemical Measurements. Anal. Chem. 2007, 79, 7526


7533.
(8) Stotter, J.; Haymond, S.; Zak, J. K.; Show, Y.; Cvackova, Z.;
Swain, G. M. Optically Transparent Diamond Electrodes for UV-Vis
and IR Spectroelectrochemistry. Interface (Electrochem. Soc.) 2003, 1,
3338.
(9) Chen, L. C.; Kichambare, P. D.; Chen, K. H.; Wu, J.-J.; Yang, J.
R.; Lin, S. T. Growth of Highly Transparent Nanocrystalline Diamond
Films and Spectroscopic Study of the Growth. J. Appl. Phys. 2001, 89,
753759.
(10) Bogdanowicz, R. Characterization of Optical and Electrical
Properties of Transparent Conductive Boron-Doped Diamond Thin
Films Grown on Fused Silica. Metrol. Meas. Syst. 2014, 21, 685698.
(11) Sobaszek, M.; Siuzdak, L.; Skowronski, K.; Bogdanowicz, R.;
Plucinski, J. Optically Transparent Boron-Doped Nanocrystalline
Diamond Films for Spectroelectrochemical Measurements on Different Substrates. IOP Conf. Ser.: Mater. Sci. Eng. 2016, 104, 012024.
(12) Joseph, P. T.; Tai, N.-H.; Chen, Y.-C.; Cheng, H.-F.; Lin, I.-N.
Transparent Ultrananocrystalline Diamond Film on Quartz Substrate.
Diamond Relat. Mater. 2008, 17, 476480.
(13) Hikavyy, A.; Clauws, P.; Maes, J.; Moshchalkov, V. V.; Butler, J.
E.; Feygelson, T.; Williams, O. A.; Daenen, M.; Haenen, K. An
Investigation of Structural and Electrical Properties of Boron-Doped
and Undoped Nanocrystalline Diamond Films. Phys. Status Solidi A
2006, 203, 30213027.
(14) Gajewski, W.; Achatz, P.; Williams, O. A.; Haenen, K.; Bustarret,
E.; Stutzmann, M.; Garrido, J. A. Electronic and Optical Properties of
Boron-Doped Nanocrystalline Diamond Films. Phys. Rev. B: Condens.
Matter Mater. Phys. 2009, 79, 045206.
(15) Zhang, Y.; Asahina, S.; Yoshihara, S.; Shirakashi, T. Fabrication
and Characterization of Diamond Quartz Crystal Microbalance
Electrode. J. Electrochem. Soc. 2002, 149, H179H182.
(16) Kulesza, S. Study of the Moderate Temperature Growth Process
of Optical Quality Synthetic Diamond Films on Quartz Substrates.
Thin Solid Films 2008, 516, 49154920.
(17) Henychova, P.; Hirmanova, K.; Vrany, M. Diamond Films for
Implantable Electrodes. Acta Polytech. 2012, 52 (5), 5861.
(18) Bogdanowicz, R.; Sobaszek, M.; Ryl, J.; Gnyba, M.; Ficek, M.;
Golunski, L.; Bock, W. J.; Smietana, M.; Darowicki, K. Improved
Surface Coverage of an Optical Fiber with Nanocrystalline Diamond
by Application of Dip-Coating Seeding. Diamond Relat. Mater. 2015,
55, 5263.
(19) Looi, H. J.; Pang, L. Y. S.; Molloy, A. B.; Jones, F.; Foord, J. S.;
Jackman, R. B. An Insight into the Mechanism of Surface Conductivity
in Thin Film Diamond. Diamond Relat. Mater. 1998, 7, 550555.
(20) Maier, F.; Riedel, M.; Mantel, B.; Ristein, J.; Ley, L. Origin of
Surface Conductivity in Diamond. Phys. Rev. Lett. 2000, 85, 3472
3475.
(21) Berkun, I. High Temperature Hall Eect Measurement System
Design, Measurement and Analysis. Ph.D. Dissertation, Michigan State
University, East Lansing, MI, 2015.
(22) ASTM F76-08: Standard Test Methods for Measuring Resistivity
and Hall Coecient and Determining Hall Mobility in Single-Crystal
Semiconductors; ASTM International: West Conshohocken, PA, 2008.
(23) Granger, M. C.; Witek, M.; Xu, J.; Wang, J.; Hupert, M.; Hanks,
A.; Koppang, M. D.; Butler, J. E.; Lucazeau, G.; Mermoux, M.; Strojek,
J. W.; Swain, G. M. Standard Electrochemical Behavior of HighQuality, Boron-Doped Polycrystalline Diamond Thin-Film Electrodes.
Anal. Chem. 2000, 72, 37933804.
(24) Barrosse-Antle, L. E.; Bond, A. M.; Compton, R. G.; OMahony,
A. M.; Rogers, E. I.; Silvester, D. S. Voltammetry in Room
Temperature Ionic Liquids: Comparisons and Contrasts with
Conventional Electrochemical Solvents. Chem. - Asian J. 2010, 5,
202230.
(25) Ren, S.; Hou, Y.; Wu, W.; Liu, W. Purification of Ionic Liquids:
Sweeping Solvents by Nitrogen. J. Chem. Eng. Data 2010, 55, 5074
5077.

6. Quasi-reversible electron-transfer kinetics was observed


for all of the redox systems at 100500 mV s1. The
semimetallic electrode material possesses a high density
of electronic states over the entire potential range
covered by the redox analytes (ca. 2 V). All of the redox
systems exhibited diusion-controlled behavior at the
BDD OTE without any conventional pretreatment. The
BDD OTE exhibits a high level of electrochemical
activity.
7. The diamond OTE is characterized by a low double-layer
capacitance in both aqueous and ionic liquid media.
Capacitance values of 27 F cm2 in H2SO4 and 0.71
F cm2 in [BMIM][BF4] were recorded. The 56 times
lower capacitance at all potentials in the RTIL is
consistent with the 5.5 times lower dielectric constant.

AUTHOR INFORMATION

Corresponding Author

*E-mail: swain@chemistry.msu.edu. Tel.: 1-517-355-1090. Fax:


1-517-353-1793.
Notes

The authors declare no competing nancial interest.

ACKNOWLEDGMENTS
This research was carried out in the framework of Project
W911NF-12-R-0011 funded by the Army Research Oce
(G.M.S.). R.J. thanks the Grant Agency of the Czech Republic
(Project P206/12/G151) and Charles University in Prague
(SVV260205) for their nancial support. This research was also
made possible by a SWE scholarship to N.W. from the Brazilian
funding agency CAPES (149/2012; AUXPE 499/2013).
R.C.R.-F. gratefully acknowledges a research productivity
scholarship from the Brazilian funding agency CNPq
(308328/2013-2). The Raman inVia Raman instrument was
acquired through the Defense University Research Instrumentation Program through Grant 65086-CH-RIP from the Army
Research Oce. We acknowledge Dr. Jonathan Counsell of
Kratos Analytical Ltd. for acquiring the XPS results using the
AXIS Supra XPS instrument.

REFERENCES

(1) Zak, J. K.; Butler, J. E.; Swain, G. M. Diamond Optically


Transparent Electrodes: Demonstration of Concept with Ferriferrocyanide and Methyl Viologen. Anal. Chem. 2001, 73, 908914.
(2) Martin, H. B.; Morrison, P. W. Application of a Diamond Thin
Film as a Transparent Electrode for In Situ Infrared Spectroelectrochemistry. Electrochem. Solid-State Lett. 2001, 4, E1720.
(3) Haymond, S.; Zak, J. K.; Show, Y.; Butler, J. E.; Babcock, G. T.;
Swain, G. M. Spectroelectrochemical Responsiveness of a Freestanding, Boron-Doped, Optically Transparent Electrode Toward
Ferrocene. Anal. Chim. Acta 2003, 500, 137144.
(4) Stotter, J.; Zak, J.; Behler, Z.; Show, Y.; Swain, G. M. Optical and
Electrochemical Properties of Optically Transparent, Boron-Doped
Diamond Thin Films on Quartz. Anal. Chem. 2002, 74, 59245930.
(5) Stotter, J.; Show, Y.; Wang, S.; Swain, G. M. Comparison of the
Electrical, Optical and Electrochemical Properties of Diamond and
Indium Tin Oxide Thin-Film Electrodes. Chem. Mater. 2005, 17,
48804888.
(6) Dai, Y.; Zheng, Y.; Swain, G. M.; Proshlyakov, D. A. Equilibrium
and Kinetic Behavior of Fe(CN)63/4 and Cytochrome c in Direct
Electrochemistry Using a Film Electrode Thin-Layer Transmission
Cell. Anal. Chem. 2011, 83, 542548.
(7) Dai, Y.; Proshlyakov, D. A.; Zak, J. K.; Swain, G. M. Optically
Transparent Diamond Electrode for Use in IR Transmission
L

DOI: 10.1021/acsami.6b02467
ACS Appl. Mater. Interfaces XXXX, XXX, XXXXXX

Forum Article

ACS Applied Materials & Interfaces


(26) Jarosova, R.; Swain, G. M. Rapid Preparation of Room
Temperature Ionic Liquids with Low Water Content as Characterized
with a ta-C:N Electrode. J. Electrochem. Soc. 2015, 162, H507H511.
(27) Heineman, W. R. Spectroelectrochemistry, Combination of
Optical and Electrochemical Techniques for Studies of Redox
Chemistry. Anal. Chem. 1978, 50, 390A402A.
(28) Szirmai, P.; Pichler, T.; Williams, O. A.; Mandal, S.; Bauerle, C.;
Simon, F. A Detailed Analysis of the Raman Spectra in Superconducting Boron-Doped Nanocrystalline Diamond. Phys. Status Solidi
B 2012, 249, 26562659.
(29) May, P. W.; Ludlow, W. J.; Hannaway, M.; Heard, P. J.; Smith, J.
A.; Rosser, K. N. Raman and Conductivity Studies of Boron-Doped
Microcrystalline Diamond, Facetted Nanocrystalline Diamond and
Cauliflower Diamond. Diamond Relat. Mater. 2008, 17, 105117.
(30) Bourgeois, E.; Bustarret, E.; Achatz, P.; Omnes, F.; Blase, X.
Impurity Dimers in Superconducting Diamond: Experiment and FirstPrinciples Calculations. Phys. Rev. B: Condens. Matter Mater. Phys.
2006, 74, 094509.
(31) Bernard, M.; Baron, C.; Deneuville, A. About the Origin of the
Low Wavenumber Structures of the Raman Spectra of Heavily BoronDoped Diamond Films. Diamond Relat. Mater. 2004, 13, 896899.
(32) Sidorov, V.; Ekimov, E. Superconductivity in Diamond.
Diamond Relat. Mater. 2010, 19, 351357.
(33) Goss, J. P.; Briddon, P. R. Theory of Boron Aggregates in
Diamond: First-Principles Calculations. Phys. Rev. B: Condens. Matter
Mater. Phys. 2006, 73, 085204.
(34) Ferrari, A. C.; Robertson, J. Origin of the 1150 cm1 Raman
Mode in Nanocrystalline Diamond. Phys. Rev. B: Condens. Matter
Mater. Phys. 2001, 63, 121405.
(35) Gonon, P.; Gheeraert, E.; Deneuville, A.; Fontaine, F.; Abello,
L.; Lucazeau, G. Characterization of Heavily B-Doped Polycrystalline
Diamond Films Using Raman Spectroscopy and Electron Spin
Resonance. J. Appl. Phys. 1995, 78, 70597062.
(36) Deneuville, A.; Baron, C.; Ghodbane, S.; Agnes, C. Highly and
Heavily Boron-Doped Diamond Films. Diamond Relat. Mater. 2007,
16, 915920.
(37) Proctor, A.; Sherwood, P. M. A. X-ray Photoelectron
Spectroscopic Studies of Carbon Fiber Surfaces. I. Carbon Fiber
Spectra and the Effects of Heat Treatment. J. Electron Spectrosc. Relat.
Phenom. 1982, 27, 3956.
(38) Proctor, A.; Sherwood, P. M. A. X-ray Photoelectron
Spectroscopic Studies of Carbon Fiber Surfaces. II. The Effect of
Electrochemical Treatment. Carbon 1983, 21, 5359.
(39) Kozlowski, C.; Sherwood, P. M. A. X-ray Photoelectron
Spectroscopic Studies of Carbon Fiber Surfaces. IV. The Effect of
Electrochemical Treatment in Nitric Acid. J. Chem. Soc., Faraday Trans.
1 1984, 80, 20992107.
(40) Hutton, L. A.; Iacobini, J. G.; Bitziou, E.; Channon, R. B.;
Newton, M. E.; Macpherson, J. V. Examination of Factors Affecting
the Electrochemical Performance of Oxygen-Terminated Polycrystalline Boron-Doped Diamond Electrodes. Anal. Chem. 2013, 85, 7230
7240.
(41) Ling, H.; Wu, J. D.; Sun, J.; Shi, W.; Ying, Z. F.; Li, F. M.
Electron Cyclotron Resonance Plasma-Assisted Pulsed Laser Deposition of Boron Carbon Nitride Films. Diamond Relat. Mater. 2002, 11,
16231628.
(42) Genisel, M. F.; Uddin, M. N.; Say, K.; Kulakci, M.; Turan, R.;
Gulseren, O.; Bengu, E. Bias in Bonding Behavior Around Boron,
Carbon and Nitrogen Atoms in Ion Implanted -BN, -BC and
Diamond-Like Carbon Films. J. Appl. Phys. 2011, 110, 074906.
(43) Hu, I.-F.; Karweik, D. H.; Kuwana, T. Activation and
Deactivation of Glassy Carbon Electrodes. J. Electroanal. Chem.
Interfacial Electrochem. 1985, 188, 5972.
(44) McCreery, R. L. Carbon Electrodes: Structural Eects on
Electron Transfer Kinetics. In Electroanalytical Chemistry: A Series of
Advances; Bard, A. J., Ed.; CRC Press: Boca Raton, FL, 1991; Vol. 17,
pp 221374.

(45) Kneten, K. R.; McCreery, R. L. Effects of Redox System


Structure on Electron Transfer Kinetics at Ordered Graphite and
Glassy Carbon Electrodes. Anal. Chem. 1992, 64, 25182524.
(46) Chen, P.; Fryling, M. A.; McCreery, R. L. Electron Transfer
Kinetics at Modified Carbon Electrode Surfaces:The Role of Specific
Surface Sites. Anal. Chem. 1995, 67, 31153122.
(47) Chen, P.; McCreery, R. L. Control of Electron Transfer Kinetics
at Glassy Carbon Electrodes by Specific Surface Modification. Anal.
Chem. 1996, 68, 39583965.
(48) Fischer, A.; Show, Y.; Swain, G. M. Electrochemical Performance of Diamond Thin-Film Electrodes from Different Commercial
Sources. Anal. Chem. 2004, 76, 25532560.
(49) Patten, H. V.; Meadows, K. E.; Hutton, L. A.; Iacobini, J. G.;
Battistel, D.; McKelvey, K.; Colburn, A. W.; Newton, M. E.;
Macpherson, J. V.; Unwin, P. R. Electrochemical Mapping Reveals
Direct Correlation between Heterogeneous Electron-Transfer Kinetics
and Local Density of States of Diamond Electrodes. Angew. Chem., Int.
Ed. 2012, 51, 70027006.
(50) Granger, M. C.; Swain, G. M. The Influence of Surface
Interactions on Reversibility of Ferri/Ferrocyanide at Boron-Doped
Diamond Thin-Film Electrodes. J. Electrochem. Soc. 1999, 146, 4551
4558.
(51) Duo, I.; Levy-Clement, C.; Fujishima, A.; Comninellis, C.
Electron Transfer Kinetics at Boron-Doped Diamond. Part I. Influence
of Anodic Treatment. J. Appl. Electrochem. 2004, 34, 935943.
(52) Macpherson, J. V. A Practical Guide to Using Boron-Doped
Diamond in Electrochemical Research. Phys. Chem. Chem. Phys. 2015,
17, 29352949.
(53) Swain, G. M.; Ramesham, R. The Electrochemical Activity of
Boron-Doped Polycrystalline Diamond Thin-Film Electrodes. Anal.
Chem. 1993, 65, 345351.
(54) Xu, J.; Chen, Q.; Swain, G. M. Anthraquinonedisulfonate
Electrochemistry: A Comparison of Glassy Carbon, Hydrogenated
Glassy Carbon, Highly Oriented Pyrolytic Graphite and Diamond
Electrodes. Anal. Chem. 1998, 70, 31463154.
(55) Chen, Q. Y.; Swain, G. M. Structural Characterization,
Electrochemical Reactivity, and Response Stability of Hydrogenated
Glassy Carbon Electrodes. Langmuir 1998, 14, 70177026.
(56) Lockett, V.; Horne, M.; Sedev, R.; Rodopoulos, T.; Ralston, J.
Differential Capacitance of the Double Layer at the Electrode/Ionic
Liquid Interface. Phys. Chem. Chem. Phys. 2010, 12, 1249912512.
(57) Williams, O. A.; Douheret, O.; Daenen, M.; Haenen, K.; O sawa,
E.; Takahashi, M. Enhanced Diamond Nucleation on Monodispersed
Nanocrystalline Diamond. Chem. Phys. Lett. 2007, 445, 255258.
(58) Williams, O. A. Nanocrystalline Diamond. Diamond Relat.
Mater. 2011, 20, 621640.
(59) Butler, J. E.; Sumant, A. The CVD of Nanodiamond Materials.
Chem. Vap. Deposition 2008, 14, 145160.
(60) Wada, N.; Solin, S. A. Raman Efficiency Measurements of
Graphite. Physica B+C 1981, 105, 353356.
(61) Mort, J.; Okumura, K.; Machonkin, M. Charge Transport in
Boron-Doped Diamond Thin Films. Philos. Mag. B 1991, 63, 1031
1036.
(62) Kornyshev, A. A. Double-Layer in Ionic Liquids: Paradigm
Change. J. Phys. Chem. B 2007, 111, 5545.
(63) Cannes, C.; Cachet, H.; Debiemme-Chouvy, C.; Deslouis, C.;
de Sanoit, J.; Le Naour, C.; Zinovyeva, V. A. Double Layer at
[BuMeIm][Tf2N] Ionic Liquid Pt or C Material Interfaces. J.
Phys. Chem. C 2013, 117, 2291522925.

DOI: 10.1021/acsami.6b02467
ACS Appl. Mater. Interfaces XXXX, XXX, XXXXXX

Vous aimerez peut-être aussi